Western University Scholarship@Western

Electronic Thesis and Dissertation Repository

11-23-2018 2:30 PM

Improved techniques for atmospheric retrievals from measurements using the Optimal Estimation Method and Machine Learning

Ghazal Farhani The University of Western Ontario

Supervisor Dr. Robert Sica The University of Western Ontario

Graduate Program in Physics A thesis submitted in partial fulfillment of the equirr ements for the degree in Doctor of Philosophy © Ghazal Farhani 2018

Follow this and additional works at: https://ir.lib.uwo.ca/etd

Recommended Citation Farhani, Ghazal, "Improved techniques for atmospheric ozone retrievals from lidar measurements using the Optimal Estimation Method and Machine Learning" (2018). Electronic Thesis and Dissertation Repository. 6020. https://ir.lib.uwo.ca/etd/6020

This Dissertation/Thesis is brought to you for free and open access by Scholarship@Western. It has been accepted for inclusion in Electronic Thesis and Dissertation Repository by an authorized administrator of Scholarship@Western. For more information, please contact [email protected]. Abstract

A new first-principle Optimal Estimation Method (OEM) to retrieve ozone number density profiles in both the troposphere and stratosphere using Differential Absorption Lidar (DIAL) measurements obtained at the Observatoire de Haute Provence (OHP) in France is described. The method is robust and applicable to any DIAL ozone lidar. The ozone retrievals are com- pared to ozonesonde measurements, and these comparisons show the profiles match within the measurement uncertainties. The OEM retrieval also successfully catches much of the structure seen by the ozonesondes. The OEM retrievals are compared with the traditional analysis, and for most heights the difference between the two methods is small. One main advantage of the OEM is that all available measurements from multiple channels as well as are used in the retrieval, eliminating the need to merge or perform corrections on the raw measurement. Thus, the tropospheric and stratospheric lidar measurements can be used together to generate an ozone profile which extends from 2.5 km to about 42 km. The upper troposphere and the lower stratosphere (UTLS) coincides with the measurements overlapping region. In the UTLS, even small changes in the distribution of the greenhouse gases can result in large changes in the atmospheric radiative forcing. The OEM can significantly improve the our understanding of the UTLS by providing an ozone density profile with a well-defined statistical and systematic uncertainty budget in this region. A new state-of-the-art machine learning technique was developed to automatically classify raw (level 0) lidar measurements to remove bad scans, and to distinguish between clear sky measurements and measurements with traces of either or . We have examined different supervised learning methods and found the random forest classifier, the support vector machine (SVM), and the gradient boosting trees could successfully classify our lidar data with more than 90% accuracy score with the random forest classifier recommended because of its greater computational speed.

Keywords: Optimal Estimation Method, DIAL, Ozone Retrievals, UTLS ozone, Machine Learning, random forest classifier, support vector machine (SVM), supervised learning, strato- sphere, troposphere, lidar

i Co-Authorship Statement

The entire thesis is written under the supervision of Dr. Robert Sica. The work presented in Chapter 2 was done in collaboration with Dr. Sophie Godin- Beekmann and Dr. Alexander Haefele. Dr. Godin-Beekmann provided me with the OHP stratospheric measurements. Also, she performed the traditional analysis of the ozone den- sity profiles. Dr. Alexander Haefele helped me to understand the Optimal Estimation Method (OEM). I wrote all the necessary MATLAB codes for the OEM retrievals. I also conducted the OEM analysis and provided comparisons between the methods. The work done in Chapter 3 is a result of collaboration with Dr. Godin Beekmann and Dr. Gerard` Ancellet who provided me with the stratospheric and tropospheric measurements as well as the traditional analysis. I was responsible for providing the MATLAB codes for the OEM retrievals, and I performed the OEM analysis. The work done in Chapter 4 is in collaboration with Dr. Alexander Haefele who provided the raw lidar measurements, and Dakota Cecil who helped me with the data preparation. Dr. Daley also helped me to understand the Machine Learning better.

ii Acknowlegements

I want to thank my supervisor Dr. Robert Sica. He trusted me and gave me all the freedom to do my research and to explore the different fields of physics. He was always patient with me, and he ever listened to my ideas and helped me to implement them in my work. I would also like to thanks Dr. Sophie Godin-Beekmann who helped me a lot with my research. For two summers I was lucky to go to France and work with her. I also thank Patricia Sica, she kindly edited my works, and we had terrific chats with each other. I will never forget our adventure in the Swiss Alps. I am grateful toward Dr. Alexey Tikhomirov, Dr. Emily McCullough, Dr. Pierre Fogal, Peter McGovern and all the staff at the Environment Canada in Eureka, Nunavut. I honestly have an amazing and unique memory of Eureka because of them! I was lucky to have the best colleagues in this department. Shayamila Mahagammulla Gamage was a fantastic colleague and a reliable friend during all these years. Robin Wing is also an incredible coworker and friend who made my stay in Paris so memorable. The staff at the Physics department are always helpful and supportive. I am genuinely appreciative toward Jodi Guthrie. From the very first day that I got to Canada, she helped me and several times she listened to me. I thank my friends Kendra, Amanda, and Moh who made my Ph.D. studies easier and made my life fabulous! Behafarid was my first roommate in London, and she became the best friend of mine. I always had her support, and I am so grateful for that. I want to thank Sina who is always there when I need him, and who listens to me for hours. My bird Coco is not a home pet to me; he is the most fantastic friend of mine. Finally, I would like to thank my parents and my sister. My parents are the reason that I started my Ph.D. During all these years, they always unconditionally supported me and stayed by my side. My sister is the one who always has my back, and I know no matter what I do, I still can go to her at any time!

iii Contents

Abstract i

Acknowlegements iii

List of Figures vii

List of Tables xvi

1 Introduction 1 1.1 Overview ...... 1 1.2 Atmospheric Structure ...... 2 1.3 Ozone in the Stratosphere ...... 3 1.3.1 Odd Catalytic Cycles ...... 5 1.3.2 Odd Chlorine Catalytic Cycles ...... 5 1.3.3 The hydroxyl radical Catalytic Cycles ...... 6 1.4 Stratospheric Ozone Depletion ...... 6 1.4.1 Ozone Trends ...... 9 1.5 Ozone in the Troposphere ...... 9 1.6 Measurements of Atmospheric Ozone ...... 11 1.7 DIAL Measurements and Retrieval of Ozone Density ...... 14 1.7.1 Ozone Density Retrievals ...... 17 1.8 Corrections Applied to the Raw Counts ...... 19 1.8.1 DeadTime Correction ...... 19 1.8.2 Signal Induced Noise ...... 19 1.8.3 Merging Process ...... 20

iv 1.8.4 Digital Filters ...... 21 1.9 Optimal Estimation Method ...... 21 1.10 Machine Learning Applications to Lidar Measurements ...... 25 1.10.1 Machine Learning Methods ...... 26

List of Appendices 1

2 Implementing the OEM to Retrieve Stratospheric Ozone Density 34 2.1 Overview ...... 34 2.2 Ozone Retrievals: Traditional versus the OEM ...... 35 2.2.1 Applying the optimal estimation method to ozone retrievals ...... 36 2.2.2 Ozone DIAL Forward Model ...... 37 2.3 Implementing the optimal estimation method retrieval ...... 39 2.4 Application of the OEM to measurements from the OHP stratospheric ozone lidar ...... 41 2.4.1 Applying the OEM to OHP measurements on 26 July 2017 ...... 43 2.4.2 Further examples of the OEM retrieval method ...... 51 2.5 Conclusion ...... 52

3 Improved ozone UTLS DIAL measurements 58 3.1 Introduction ...... 58 3.2 METHODOLOGY ...... 61 3.2.1 Implementing the OEM for the OHP lidars ...... 63 3.2.2 Description of the Lidars ...... 67 3.3 OEM ozone retrieval in the free troposphere and stratosphere ...... 68 3.4 Tropospheric Ozone Lidar Retrievals ...... 76 3.5 Summary ...... 79 3.6 Conclusions ...... 80 3.6.1 Dataset Citation ...... 81

4 Machine Learning Methods in Lidar Measurements Classification 86 4.1 Introduction ...... 86

v 4.2 Instrumentation ...... 87 4.3 A Brief Introduction to Machine Learning ...... 89 4.3.1 Supervised Approach ...... 89 4.3.2 Unsupervised Approach ...... 91 4.4 Learning Algorithms ...... 92 4.4.1 Support Vector Machine Algorithms ...... 93 4.4.2 Decision Trees Algorithms ...... 93 4.4.3 Random Forests ...... 96 4.4.4 Gradient Boosting Tree Methods ...... 97 4.4.5 The t-distributed Stochastic Neighbour Embedding Method ...... 98 4.5 Result for supervised and unsupervised learning in PCL and RALMO system . 100 4.5.1 Unsupervised result ...... 102 4.6 Summary ...... 104 4.7 Conclusion ...... 106

5 Conclusions and Future Work 109

Curriculum Vitae 111

vi List of Figures

1.1 The mean atmospheric profile is plotted for the month of July at the Observatoire de Haute-Provence 44◦N, 5.8◦E. The data is retrieved from the Mass Spectrometer Incoherent Scatter Radar (MSIS) database (Hedin, 1991). 4

1.2 In the DIAL technique two wavelengths (the “on-line” and the “off-line” wave- lengths) are simultaneously transmitted to the atmosphere. The back scattered signals are collected by a large mirror or by multiple small mirrors. The col- lected signal (via optical fibers) are sent to the PMTs...... 15

1.3 The flowchart is the summery of the steps we take to calculate the retrieval and its associated uncertainties...... 25

2.1 To implement the OEM, the a priori profiles for ozone and air density, back- ground counts, dead time values, and lidar constants are needed. Moreover, b parameters should be identified and proper values for them should be cal- culated. The covariance matrices for a priori profiles, measurements, and b parameters need to be calculated as well...... 39

2.2 Average count rates for 5 hours of measurements on 26 July 2017. Left panel:“on- line” wavelength (blue curve, low altitude; red curve, high altitude). Right panel: “off-line” wavelength (blue curve, low altitude ; red curve, high altitude). 44

vii 2.3 Averaging kernels for the ozone density for the measurements on 26 July 2017. The horizontal dashed line is a height below which the OEM retrievals is more than 80% due to the measurements. Above this horizontal cut-off as the SNR drops, the retrieval starts to fall back to the a priori profile. For clarity, the averaging kernels are only shown every 1500 m in altitude. The red line shows the summation of rows in the averaging kernel matrix at each altitude. The summation is of order unity below 42.7 km...... 45

2.4 Residuals between the forward model and the measurements for the “on-line” and “off-line” channel (blue curves). The red line shows the uncertainty of the measurements...... 46

2.5 Right panel: The statistical uncertainty of the OEM with correlation lengths

(`s = 1000 and 1400 m is plotted (red curve) against the statistical uncertainty

of the OEM with correlation lengths `s = 1400 m and 5500 m (black dotted curve). Additionally, the uncertainty of retrieval in the traditional method (blue curve) is plotted. The retrieval uncertainties in the OEM and the traditional method can be compared. Left panel: The vertical resolution of the OEM with

correlation lengths (`s = 1000 and 1400 m (red curve) is plotted against the

vertical resolution of the OEM with correlation lengths `s = 1400 and 5500 m (black dotted curve). The vertical resolution of the traditional method is shown as well (blue curve). The horizontal dashed line indicated the maximum height at which the retrieval is independent from the a priori...... 47

2.6 At height from 20 km to 40 km, the uncertainty of retrieval for the traditional method (assuming that it has a vertical resolution similar to the OEM vertical resolution) is plotted against the OEM retrieval uncertainty (blue curve: OEM; red curve: traditional). The horizontal dashed line indicated the maximum height at which the OEM retrieval is independent from the a priori...... 48

viii 2.7 OEM ozone retrieval (red curve) from 20:07 UT to 00:15 UT on 26 July 2017 as well as the ozonesonde profile (green curve) and the traditional ozone retrieval (blue curve) are plotted. The dashed black line shows the OEM retrieval when

the correlation length (łS ) became larger. The horizontal dashed line shows the cut-off below which the effect of the a priori ozone profile is small less tan 10%. 49

2.8 For the night of 26 July 2017. Left panel: The percentage difference be- tween the OEM retrieval and the ozonesonde measurements in the form of:

OEM−sonde ( sonde ∗100) (blue curve); the percentage difference between the traditional traditional−sonde retrieval and the ozonesonde measurements in the form of ( sonde ∗ 100) (red curve). Right panel: The percentage difference between the OEM retrieval and the traditional retrieval (blue curve); the summation of the statistical un- certainty of the traditional and OEM retrievals (red curve)...... 50

2.9 For the night of 26 July 2017. The left panel: the OEM retrieval using the US standard model as a priori profile (purple curve) and the OEM retrieval using the OHP climatology as a priori profile (red curve) are plotted. Furthermore, the traditional method retrieval (blue curve) is plotted, thus the OEM retrievals can be compared with each other and with the traditional retrieval. The right panel: Percentage difference between the OEM retrievals using the two dif- ferent a priori profiles (blue curve) is plotted. This difference is with in the retrieval uncertainty. At higher altitudes (above 35 km), when the SNR drops, the difference between the two methods is less than 5%, which is smaller than the retrieval uncertainty at that height...... 51

2.10 For the night of 26 July 2017. The statistical uncertainty of the OEM (blue), the Rayleigh-scatter cross section uncertainty at 308 nm (red), the ozone absorption cross section at 308 nm (orange), and the ozone absorption cross section for the 355 nm channel (purple). The horizontal dashed line shows the height below which the retrieval is independent of the a priori profile...... 52

ix 2.11 Left panel: The retrieved air density (blue line) is plotted against the a priori profile (red line). Mid panel: The percentage difference between the scaled relative air density generated from the Raman channel and the OEM air density retrievals. The difference is less than 10%. Right panel: The statistical un- certainty of the OEM retrieval of air density (blue), the Rayleigh-scatter cross section uncertainty for the 308 nm channel (red), and the ozone absorption cross section in both channels (purple)...... 53

2.12 Left Panel: OEM ozone retrieval on the night of 14 July 2017 (red curve) compared to the ozonesonde profile (green curve) and the traditional ozone re- trieval (blue curve). Right Panel: OEM ozone retrieval on the night of 20 July 2017 (red curve) compared to the ozonesonde profile (green curve) and the traditional ozone retrieval (blue curve). These cases demonstrate the high res- olution of the OEM technique as evidenced by the excellent agreement around the ozone peak with the sonde measurement...... 53

2.13 For the night of 14 July 2017, (a) the percentage difference between the tradi- tional method and the OEM retrieval (blue curve) plotted within the envelope of the total statistical uncertainty of the two method (red curve). The agreement between the two lidar ozone determinations are within the statistical uncer- tainty above 17 km. (b): The red curve is the percentage difference between the OEM retrieval and sonde measurements. The blue curve is the percentage difference between the traditional method and sonde measurements. Figures (c) and (d) are the same format as (a) and (b) for the night of 20 July 2017. . . . 54

3.1 Averaging kernels for tropospheric-stratospheric ozone measurements. The av- eraging kernels are only shown every 450 m for lower altitude (from 2.5 km to 11 km height), and every 1500 m in higher altitudes for clarity. As is shown by the response function (red curve), the area of the averaging kernel matrix has a small spike at 11 km, when the stratospheric ozone measurements are included. 70

x 3.2 The percentage difference between the forward model and the actual measure- ments are shown in blue. The statistical uncertainty is plotted in red. The four plots at the left are stratospheric forward model residuals, and the four plots at the right are the tropospheric forward model residuals. For digital counting systems, the Poisson distribution is appropriate and the variance of the mea- surements at each altitude is the number of photons at that altitude. However, the output signals of the analog channels do not follow a Poisson distribution and to find the variance a 3-point running filter is used. As a result the red line (which indicates the noise of measurements) for the analog channels is more structured than for the digital channels...... 70

3.3 The OEM retrieval (red curve) compared to the traditional calculation of ozone using the OHP lidar systems. The tropospheric lidar starts at 2.5 km and ex- tends upward to 14.5 km (blue curve). The ozone profile measured by the stratospheric lidar system (black curve) overlaps with the profile retrieved from the tropospheric lidar system in the UTLS. In this region, the OEM retrievals smoothly transition from relying primarily on the tropospheric lidar measure- ments to the stratospheric measurements...... 72

3.4 Left panel: The statistical uncertainty of the OEM retrieval (red curve) is plot- ted against the statistical uncertainty of the traditional retrievals. The uncer- tainty of the retrieval for the stratospheric and tropospheric lidar systems re- spectively are shown in black and blue. Right panel: The vertical resolution of the OEM retrieval is shown in red. The vertical resolution of the traditional cal- culation from the tropospheric lidar system is shown in blue, while the vertical resolution of the retrieved profile produced from the stratospheric lidar system is shown in black...... 73

3.5 The percentage difference between the OEM retrieval and the ozonesonde mea- surements (blue curve) is plotted within the total statistical uncertainty of the OEM retrievals plus the ozonesonde measurement (red curves)...... 73

xi 3.6 Uncertainty budget on the night of 12 July 2017. The statistical uncertainty of the retrieval (blue), the Rayleigh-scatter cross section uncertainty at 308 nm (dashed line red), the Rayleigh-scatter cross section uncertainty at 289 nm (dashed line yellow), the ozone absorption cross section at 308 nm (dashed line purple), the ozone absorption cross section for the 289 nm channel (dashed line green), and the overlap function for the 289 nm channel (dashed line blue) all contribute to the budget. The horizontal dashed line shows the height below which the retrieval is independent of the a priori profile...... 74

3.7 OEM ozone-profile retrieval on 14 July 2017 (red curve). The tropospheric tra- ditional retrieval (blue curve) extends from 2.5 km to 15 km, while the strato- spheric traditional retrieval (black curve) extends from 10 km to 43 km. At the region where the tropospheric and stratospheric lidar measurements overlap, the OEM can smoothly makes a transition from one lidar system’s measure- ments to the other system’s measurements...... 75

3.8 OEM ozone-profile retrieval on 26 July 2017 (red curve). The tropospheric tra- ditional retrieval (blue curve) extends from 2.5 km to 14 km, while the strato- spheric traditional retrieval (black curve) extends from 12.5 km to 43 km. At the region where the tropospheric and stratospheric lidar measurements over- lap, the OEM can smoothly makes a transition from one lidar system’s mea- surements to the other system’s measurements...... 76

3.9 Averaging kernels for the ozone density for the measurements on 12 July 2017. The horizontal dashed line is the height cut-off above which the sensitivity of the retrieval to measurements is less than 90%. The averaging kernels are only shown every 450 m in altitude. The summation of rows in the averaging kernel matrix, for each specific height, is shown by the red curve...... 77

3.10 Both the OEM retrieval (red curve) and the traditional retrieval (blue curve) extend from 2.5 km to 14.2 km. The ozonesonde profile is plotted in green. The black dashed line defines the cut-off altitude of the retrieval...... 77

xii 3.11 Left panel: The statistical uncertainty of the OEM retrieval (red curve) as well as the statistical uncertainty of the traditional retrieval (blue curve) for 12 July 2017. Right panel: Vertical resolution of the OEM retrieval is shown in red, while the vertical resolution of the traditional retrieval is shown in blue. The spike at 5.5 km in the OEM uncertainty is from the inclusion of the digital channels at this height...... 78

3.12 Comparison of tropospheric OEM ozone DIAL retrievals for the nights of 14 July and 26 July 2017 (red curves), the traditional method (blue curves), and ozonesonde measurements (green curves). The horizontal dashed line is the altitude below which the OEM retrieval is mostly independent of the a priori ozone profile assumed. Left panel: 14 July 2017; right panel: 26 July 2017. . . 79

4.1 The training set is a matrix of size (m ∗ n), in which each row represents one dataset (here a lidar scan), and each element in each column (except for the n-th column) represents an attribute (for lidar scans: photon counts at each altitude). The n-th column holds the label for each scan (here we have four different labels: 0, 1, 2, and 3)...... 90

4.2 Example of scans taken by PCL for Rayleigh and Raman channels. Panel (a) shows examples of bad scans. In this plot, the signals in blue and dark red have extremely low power, the purple signal has extremely high background counts, and the signal in orange has a distorted shape and high background counts. Panel (b) shows two examples of good signals: the signal in red has a high laser power compared to the blue curve. In the supervised approach, the red signal is labeled 3, and the blue signal is labeled 1. Panel (c) shows a scan in which clouds occur at lower altitudes...... 91

4.3 Using a t-SNE unsupervised algorithm, the iris dataset is clustered into three different types. Setosa is shown in violet, versicolor is shown in blue and vir- ginica is shown in yellow...... 92

xiii 4.4 Example of linear SVM: the hyperplane separates sun and star logos. The subset of data points within the margin (support vector) is the most important data point as they play an important role in defining the optimum hyperplane. IN SVM a wider margin is preferable, as it divides two classes. This figure is adapted from: (Burges, 1998)...... 94

4.5 The schematics of a decision tree is shown. From the root (a node with no parents), a decision question is asked and answered in the form of yes (1) or no (0). The answers (yes or no) generate two children. Leaves are the nodes with no children. Ideally, each leaf is pure (all the points inside it has the same classification)...... 95

4.6 The schematic of a random forest: each tree casts a vote; based on majority vote, a data point is classified...... 97

4.7 Boosting methods are based on the idea that combining many “weak” approxi- mation models (a learning algorithm which is Using the t-SNE algorithm scans for the low-gain Rayleigh measurement channel on the night of May 15 2012 were clustered to four different groups. As shown in the figure, cluster number 2 has some outliers...... 98

4.8 Red curve: the Gaussian distribution for data points (extending from -5 to 5 in x-axis). The peak of the distribution is at its mean in 0. Blue curve: The Student’s t-distribution for the same data points. The distribution is heavy- tailed (comparing to the Gaussian distribution)...... 99

4.9 Using the t-SNE algorithm scans for the low-gain Rayleigh measurement chan- nel on the night of May 15 2012 were clustered to four different groups. As shown in the figure, the cluster number 2 has some outliers...... 103

xiv 4.10 Left Panel: All scans collected for the PCL system from the low-gain Rayleigh channel which were plotted for the night of May 15 2012. The green signals have extremely low power. The red line represents all signals with lower laser signal and the blue line indicates the signals which are good scans. The black lines are signals with extremely high backgrounds. Right Panel: Each line rep- resents an average of the signals within a cluster. The black line indicates the outliers which have extremely high background counts and are outliers belong- ing to cluster 3 (blue curve). The red line is the average signal for scans with lower laser power. The green line is the average signal for cluster 2. The back- ground counts in the green line start at lower latitude of 40 km, where as for the red line the background starts at 60 km and in the blue line the background starts at 85 km...... 104 4.11 Figure. 4.11 (a): Using the t-SNE algorithm, scans for the Nitrogen channel on the night of May 15 2012, collected by the PCL system, were clustered into two different groups. Figure. 4.11 (b): The red line (cluster number 2) is the average of all signals within this cluster and indicates the scans in which clouds are detectable. The blue line (cluster number 2) is the average of all signals within this cluster and indicates the clear scans (non-cloudy condition). . 105 4.12 The result of applying the t-SNE algorithm to different channels at different nights is plotted. Each cluster contains lidar scans sharing similar features. . . . 105

xv List of Tables

2.1 Values and associated uncertainties for the retrieved and forward model param- eters...... 42 2.2 Dead time values which were calculated for each channel on the night of 26 July 2017...... 51

3.1 Values and associated uncertainties for the retrieved and forward model param- eters...... 67 3.2 Measurement periods for the tropospheric and stratospheric lidars systems and launch times for the ozonesondes...... 69

4.1 The success score for each of the algorithms for 2000 scans of PCL lidar. . . . . 101 4.2 Number of scans which used in each type...... 101 4.3 The success score for each of the algorithms for day time scans of the RALMO lidar...... 102 4.4 The success score for each of the algorithms for day time scans of the RALMO lidar...... 102 4.5 Number of scans used in each type; for day time and night time we used the same number of scans...... 102

xvi Chapter 1

Introduction

1.1 Overview

Ozone is a minor constituent in the atmosphere that mostly resides in the stratosphere. Strato- spheric ozone represents about 90% of the total column ozone, and about 10% of the ozone is concentrated in the troposphere. By absorbing solar ultraviolet (UV) radiation, stratospheric ozone protects the Earth’s surface from receiving excessive radiation, and this heat causes the positive temperature gradient in the stratosphere (Andrews et al., 1987). The significant decline of stratospheric ozone became a serious concern at the end of the last century. The discovery of an ozone hole in the Antarctic and the substantial reduction in the total column ozone (Farman et al., 1985; WMO, 1988) were clear examples of ozone depletion due to heterogeneous reactions involving ozone-depleting substances (ODSs). Since 1989, the implementation of the Montreal Protocol has successfully reduced the level of ODSs in both the troposphere and the stratosphere. The effect of the decline of ODSs on the recovery of the ozone layer has yet to be carefully observed and studied. Thus, continuing measurements of stratospheric ozone to characterize the rate of ozone recovery is required. In contrast to stratospheric ozone, ozone molecules residing in the upper troposphere act as the third most abundant greenhouse gas contributing to the radiative forcing of climate change (Ramaswamy et al., 2001). Furthermore, ozone at the surface is a pollutant and has an adverse effect on air quality, human health, and the ecosystem. Continuous measurements of the strato- spheric and the tropospheric ozone enable us to monitor and understand ozone changes and its

1 2 Chapter 1. Introduction trends in the atmosphere. Various airborne and ground-based instruments have been employed to measure ozone density. This thesis focuses on Light Detecting and Ranging (LIDAR) system, which is a ground-based instrument. Lidar’s high vertical and temporal resolution make it a suitable system for ozone monitoring. In Chapter 1, the importance of the ozone in the atmosphere is explained in detail. Section 1.2 is devoted to explaining the structure of the atmosphere. In Section 1.3, the role of the stratospheric ozone is discussed. In Section 1.4, a summary of ozone depletion processes and ozone changes before and after the implementation of the Montreal Protocol is presented. Section 1.5 discusses the role of the tropospheric ozone in the atmosphere, and Section 1.6 briefly describes the instruments that are typically used in ozone studies. Sections 1.7 and 1.8 describe the DIAL system and ozone retrievals in detail. In Section 1.9 we introduce the Optimal Estimation Method, and in Section 1.10 we briefly describe the importance of Machine Learning approach for lidar data classification. The focus of Chapter 2, and Chapter 3 is to introduce and to implement the Optimal Esti- mation Method (OEM) which is used to retrieve tropospheric and stratospheric ozone profiles, using a Differential Absorption Lidar (DIAL). Chapter 4 of this thesis focuses on implementing a state-of-the-art machine learning method for lidar data classification. In lidar data analysis, before applying any algorithm or any pre- processing (correcting photon counts), each lidar scan should be examined for quality purposes (when dealing with level-0 raw measurements). This quality check to remove background counts is mostly done manually, which is a time-consuming and inaccurate process. We pro- pose using machine learning techniques for raw-data classification and automating this quality check.

1.2 Atmospheric Structure

The atmosphere is a layer of gas which surrounds the Earth. Based on the vertical distribution of temperature, the atmosphere is separated into four different layers (see Fig. 1.1). The lowest layer is called the troposphere. 1.3. Ozone in the Stratosphere 3

The main characteristic of the troposphere is its negative lapse rate. The lapse rate is defined

dT as Γ = − dh and is the negative of the rate of temperature change with respect to height (negative lapse rate means that the temperature decreases with height). The troposphere is statically unstable; as a result, the tropospheric gases are well mixed. Moreover, most of the weather takes place in this layer, and it contains most of the . The tropopause is the boundary line where change in the lapse rate occurs. The tropopause is defined as the lowest level at which the rate of lapse decrease is 2 ◦C/km or less (WMO, 1992). The tropopause has a height of about 8 km at the poles and 15 km at the equator (Harrison, 2007). The second layer is called the stratosphere which extends from the tropopause to roughly 50 km in altitude. In the stratosphere, where ozone density is at its peak, ozone molecules absorb UV radiation from the Sun. Thus, in this layer, the temperature increases with height. Compared to the troposphere, the stratosphere is generally stable, very dry, and ozone rich. The boundary that separates the stratosphere from the next layer is called the stratopause. The mesosphere is located above the stratopause where the temperature decreases again. The air in this region is extremely thin, and 99.9% of the atmospheric mass is found below the mesosphere. In the upper layer of the mesosphere, because of the passage of meteors, “dust” exists. Moreover, in this layer high concentration of sodium, potassium, and iron can be observed. The abundance of these elements is related to meteor showers. The mesosphere is bound by the mesopause, which has a height of approximately 85 km. The thermosphere is located above the mesopause. A significant temperature inversion, due to the absorption of energetic solar radiation by molecules, can be observed in this layer (Ahrens, 1998).

1.3 Ozone in the Stratosphere

The bond energy of the oxygen molecule is 498 kJ mol−1, which corresponds to the energy of a photon with a wavelength of 240 nm, thus only photons with shorter wavelengths than 240 nm can photolysis oxygen molecules. As is shown in (R 1), the photolysis of an oxygen molecule produces two oxygen atoms in their ground-level triplet state O(3P). The oxygen atoms are highly re-active, and they rapidly combine with O2 to form ozone (R 2) molecules, where M can be any molecule. Ozone molecules, with a bond energy of 445 kJ mol−1, weaker than 4 Chapter 1. Introduction

Figure 1.1: The mean profile is plotted for the month of July at the Observatoire de Haute-Provence 44◦N, 5.8◦E. The data is retrieved from the Mass Spectrometer Incoherent Scatter Radar (MSIS) database (Hedin, 1991).

O2. Thus photons with lower energy (corresponding to wavelengths shorter than 270 nm) can photolyze ozone molecules (see R 3). The oxygen atoms O(1D) are in their excited singlet

3 state, but by colliding with N2 or O2 molecules they will stabilize to O( P). The rate of reaction for R 2 and R 3 is much faster than the rate of reaction for R 1 and R 4. Thus there is a rapid

cycling between the atomic oxygen and ozone. It is useful to look at O and O3 together as

an odd oxygen family distinct from O2 (even oxygen) which is a much longer-lived species. More detail about the Chapman mechanism can be found in Chapman (1930) and Brasseur and Solomon (2006).

O2 + hν −−−→ O + O (R 1)

O + O2 + M −−−→ O3 + M (R 2)

O3 + hν −−−→ O + O2 (R 3)

O3 + O −−−→ O2 + O2 (R 4) .

The Chapman cycle is based on oxygen-only chemistry; however, the stratospheric ozone 1.3. Ozone in the Stratosphere 5 is also destroyed by hydrogen and nitrogen oxide chemistry (Bates and Nicolet, 1950; Crutzen, 1970; Johnston, 1992). Each of these species has their own odd families and can destroy odd oxygen in a catalytic process. Chlorine can also play an important role in ozone destruction as it can engage in catalytic cycles with odd oxygen (Stolarski and Cicerone, 1974).

1.3.1 Odd Nitrogen Catalytic Cycles

The nitrogen family has two cycles which destroy stratospheric ozone. The first cycle, which is dominant in the middle stratosphere (Brasseur and Solomon, 2006), involves both atomic oxygen and ozone. The second cycle does not require atomic oxygen and is more critical below 30 km. Both cycles are shown below:

NO + O3 −−−→ NO2 + O2 (R 5)

O + NO2 −−−→ NO + O2 (R 6)

Net Cycle 1: O + O3 O2.

NO + O3 −−−→ NO2 + O2 (R 7)

NO2 + O3 −−−→ NO3 + O2 (R 8)

NO3 + hν −−−→ NO + O2 (R 9)

Net Cycle 2: 2 O3 3 O2.

During the day, the oxidation of NOx by OH leads to the formation of HNO3. During

the night, NO3 and NO2 can bond and form N2O5. Thus HNO3 and N2O5 are reservoirs for

NOx. Details about all of these reactions can be found in Jacob et al. (1996) and Brasseur and Solomon (2006).

1.3.2 Odd Chlorine Catalytic Cycles

In the stratosphere, due to photolysis of organic chlorine species (e.g., CFCs, HCFCs and CCl4) atomic chlorine (Cl) is produced. The free Cl reacts with chlorine monoxide radicals (ClO) as shown below:

Cl + O3 −−−→ ClO + O2 (R 10)

ClO + O −−−→ Cl + O2 (R 11) Net Cycle: O + O3 O2. 6 Chapter 1. Introduction

Atomic oxygen (which is needed for the second reaction of this cycle) is formed when UV radiation reacts with ozone and oxygen molecules. Thus, this cycle is more critical at mid-latitudes and the tropics where the UV radiation is more intense.

When Cl and ClO react with CH4 and NO2 they convert to HCl and ClONO2 reservoirs, thus the above cycle stops:

Cl + CH4 −−−→ HCl + CH3 (R 12)

ClO + NO2 + M −−−→ ClNO3 + M · (R 13)

More detail on odd chlorine catalytic cycles can be found elsewhere (Bates and Nicolet, 1950).

1.3.3 The hydroxyl radical Catalytic Cycles

Chlorine and bromine monoxides can interact hydroxyl HOx and cause ozone destruction as follows:

O3 + OH −−−→ HO2 + O2 (R 14)

XO + HO2 + −−−→ HOX + O2 (R 15) HOX + hν −−−→ X + OH (R 16)

O3 + X −−−→ XO + O2 (R 17) where X stands for either chlorine or bromine monoxides. In the lower most stratosphere, where oxygen atoms are rare, the hydroxyl chemistry dominates the gas phase loss for ozone molecules.

1.4 Stratospheric Ozone Depletion

Due to high levels of ozone-depleting substances (ODSs), chemical depletion of total ozone has been detected globally since the mid-1970s (WMO, 1999). In 1974, studies suggested that anthropogenic chlorofluorocarbons (CFCs) were major sources of stratospheric chlorine, and could play an important role in stratospheric ozone destruction (Molina and Rowland, 1974). Moreover, the gas phase reactions which were introduced earlier in this chapter could not explain the observed massive loss of the ozone. 1.4. Stratospheric Ozone Depletion 7

Ozone depletion is most pronounced in Antarctica, where the stratosphere is characterized by the presence of a strong polar vortex from May to November. During the Polar Night, the stratosphere temperature drops and a low pressure system centered over the polar region, known as the polar vortex, develops. The air within the polar vortex has very low temperature, which causes a strong temperature gradient between the mid-latitudes and high-latitudes. The strong temperature gradient results in air movement from the equator to the poles, which is directed to the east by the Coriolis effect. This situation results in the formation of large horizontal pressure gradients and high jet at the edge of the vortex, known as the polar night jet. These winds can isolate the air inside the vortex from the warmer and ozone rich air masses at lower latitudes.

Low polar temperature conditions are key to severe ozone depletion. At low , Polar Stratospheric Clouds (PSCs) in the form of liquid or solid are formed. In polar regions, different types of liquid and solid PSCs are formed when the stratospheric temperature drops below −78◦C. PSCs are formed between 12 km to 25 km in altitude and over large areas. Nitric acid, sulfuric acid and droplets of supercooled water (NAT) compose type IPSCs. As the tem- perature in the stratosphere drops below the frost point of supercooled water droplets (−88◦C) the NATs crystallize into type II PSCs.

In Spring during Polar Sunrise ODSs chlorine compounds are activated towards species that are more detrimental for the ozone in the presence of solar radiation. PSCs play a sig- nificant role in this process, as some reservoirs such as HCl and ClONO2, can react on the surface of PSCs, releasing active chlorine atoms. As a result, most of the stratospheric ozone at altitudes between 15 km to 25 km disappears (Farman et al., 1985; WMO, 2011, 2014). This phenomenon is known as the Antarctic ozone hole.

Here, one of the surface reactions is shown:

HCl + ClNO3 −−−→ Cl2 + HNO3 (R 18)

Cl2 + hν −−−→ 2Cl · (R 19)

Several studies suggested that inside the polar vortex high concentration of ClO has a major effect on the ozone’s destruction (Jacob et al., 1996). Due to extremely low temperatures, high amounts of ClO are sustained inside the polar vortex. During the spring because of the sun 8 Chapter 1. Introduction light the ClO cycle can occur:

Cl + O3 −−−→ ClO + O2 (R 20)

Cl + O3 −−−→ ClO + O2 (R 21)

ClO + ClO + M −−−→ Cl2O2 + M (R 22)

Cl2O2 + hν −−−→ Cl + ClO2 (R 23)

ClO2 + M −−−→ Cl + O2 + M (R 24) Net Cycle: 2 O3 3 O2.

Furthermore, McElroy et al. (1986) and Tung et al. (1986) showed that chlorine and bromine reactions can also destroy stratospheric ozone:

Cl + O3 −−−→ ClO + O2 (R 25)

Br + O3 −−−→ BrO + O2 (R 26)

BrO + ClO −−−→ Br + ClO2 (R 27)

ClO2 + M −−−→ Cl + O2 + M (R 28) Net Cycle: 2 O3 3 O2.

As a result of these reactions, the abundance of Cl gases will significantly increase. More- over, PSCs are the main reason for stratospheric denitrification. Below, we briefly discuss the denitrification process. Most types of PSCs form from nitric acid (HNO3) and water (which is condensed on liquid sulfuric- acid-containing particles). As PSCs contain large particles, due

to gravity, they can descend several kilometers. Thus, in the process, large amounts of HNO3 are removed from the stratosphere which is known as the denitrification of the stratosphere. In

the stratosphere, nitric acid is the source of NOx which can convert the highly reactive Cl atoms

to the reservoir molecules (ClONO2). Thus, by denitrification of the stratosphere Cl remains and destroys ozone molecules.

Due to stronger planetary wave activity in the Northern Hemisphere, the polar vortex in the Arctic is less symmetric and can be very unstable; thus ozone loss in the Arctic is generally less severe. Overall stratospheric temperature in the Arctic is also higher, and there are fewer PSCs in the Arctic. Moreover, before depletion starts, as a result of stronger transport of the ozone from the tropics to the Northern Hemisphere, the abundance of ozone molecules in the Arctic is more than in the Antarctic. However, in 2011, substantial loss of the stratospheric ozone in the Northern Hemisphere was reported in the Arctic by Manney et al. (2011). 1.5. Ozone in the Troposphere 9

1.4.1 Ozone Trends

In the early 1990s, the total amount of global ozone was about 5% less than the average amount of ozone from 1964 through 1980 (WMO, 2014). Under the Montreal Protocol and its sub- sequent amendments, the emission and thus abundance of anthropogenic ODSs in the tropo- sphere has decreased from its peak in 1994 by approximately 10% (WMO, 2014). As a result, the amount of ozone during the early 2010s lessened to 3% less than the average amount of ozone in the 1964-1985 period. Recently, the first signs of total ozone recovery over Antarc- tica was observed (Solomon et al., 2016). However, for non-polar regions, since 2000, no significant positive trend is detected (WMO, 2014). Although the trends in the total column ozone are insignificant, in the upper stratosphere (around 42 km) the ozone level has significantly increased (Harris et al., 2015). This increase does not indicate that ozone in the whole stratosphere is increasing. In contrast, many studies have suggested that, at mid-latitudes and tropical latitudes, the ozone content, at the lower stratosphere has continued to decrease (Ball et al., 2018). Trends in the ozone are on the order of few percent, for example, trends in the upper stratosphere are around 1% to 3% per decade (Harris et al., 2015). Thus, it is crucial to take ozone measurements with an instrument with high spatial and temporal resolution to detect these changes.

1.5 Ozone in the Troposphere

A small amount of ozone resides in the troposphere, where it is a greenhouse gas contributing to climate change. Moreover, tropospheric ozone near ground-level is an air pollutant damaging human health and threatening ecosystem health. The tropospheric ozone budget depends on both photochemical and physical processes. Because of the high concentration of ozone in the stratosphere, it was once assumed that ozone transportation from the stratosphere to the tropo- sphere is the dominant source of tropospheric ozone (Junge, 1962; Danielsen, 1968). However, later studies showed that the tropospheric ozone is mostly produced from the photochemical oxidation of CO and hydrocarbons (catalyzed by HOx and NOx). Here, we briefly describe the two processes. 10 Chapter 1. Introduction

• Stratosphere-Troposphere Exchange (STE) In the mid-latitudes and the polar regions, air masses move along lines which have a con- stant (isentropic lines), as a result, in the UTLS region, air masses from the ozone-rich stratosphere can irreversibly move down towards the upper tropo- sphere. This process is adiabatic; thus it does not require heat. In the STE process, chem- ical constituents such as ozone molecules are depleted from the stratosphere (where their abundance is necessary), and their concentration increases in the troposphere (where they are greenhouse gases) (Holton et al., 1995). The rate of this exchange is between 770 ± 400 Tg/year (IPCC2007). Depending on the latitude, tropopause height, and season, the contribution of STE to tropospheric ozone concentration can vary. During spring, at high latitudes and 500 hPa, the stratospheric contribution to the tropospheric ozone is about 40%, this value drops to 25% during fall. At mid-latitudes, during spring, the contribution is between 35% and 40%, and during fall it drops to 10-15% (Cohen et al., 2018).

• Photochemistry

Photochemistry involving CO, CH4, and Volatile Organic Compounds (VOCs), in the presence of nitric oxides, is another source of tropospheric ozone. These molecules are

known as ozone precursors. The main driver of tropospheric ozone concentration is NOx.

The primary source of NOx in the troposphere is fossil fuel combustion (Finlayson-Pitts et al., 1999). Natural sources, including soil emissions and lightning, can contribute to

NOx formation as well (Sauvage et al., 2007). However, less than one-third of NOx is produced from natural sources. The net flux of tropospheric ozone due to photochemical

activities is 3420 ± 770 Tg/year (IPCC2007). Here, the reaction involving CO, CH4, and Volatile Organic Compounds (VOCs) are discussed briefly. The contribution of CO to ozone production is given by the following reactions:

O2 CO + OH −−−→ CO2 + HO2 (R 29)

HO2 + NO −−−→ OH + NO2 (R 30)

O2 NO2 + hν −−−→ NO + O3 (R 31) Net Cycle: CO + 2 O2 CO2 + O3. 1.6. Measurements of Atmospheric Ozone 11

The reactions involving CH4 are shown below:

CH4 + OH −−−→ CH3 + H2O (R 32)

CH3 + O2 + M −−−→ CH3O2 + M (R 33)

CH3O2 + NO −−−→ CH3O + NO2 (R 34)

CH3O + O2 −−−→ CH2O + HO2 (R 35)

O2 CH2O + hν −−−→ CHO + O2 (R 36)

CHO + O2 −−−→ CO + HO2 (R 37)

O2 CO + OH −−−→ CO2 + HO2 (R 38)

4 (HO2 + NO −−−→ OH + NO2) (R 39)

5 (NO2 + hν −−−→ NO + O3) (R 40) Net Cycle: CH4 + 10 O2 CO2 + H2O + 5 O3 + 2 OH.

Finally, the VOCs contribution is as follows:

VOC1 + 4O2 + 2hν −−−→ VOC2 + H2O + 2O3 · (R 41)

Reactions involving HOx are the main source for tropospheric ozone destruction:

HO2 + O3 −−−→ OH + 2O2 (R 42)

OH + O3 −−−→ HO2 + O2 · (R 43)

The balance between the destruction and construction of the tropospheric ozone is deter-

mined by the abundance of the precursors and HOx molecules. A detailed discussion on the tropospheric ozone distribution and trends can be in Gebhardt et al. (2014).

1.6 Measurements of Atmospheric Ozone

Different instruments have been used to measure ozone concentration in both the troposphere and the stratosphere, each with various advantages and disadvantages. In the following section, a brief introduction to ozone measurements will be given. Both remote sensing and in-situ techniques have been used to measure the concentration of ozone in the atmosphere. For in-situ measurements, a sample of air is taken from the atmo- sphere and is analyzed to determine the ozone content. In remote sensing measurements, pas- 12 Chapter 1. Introduction sive and active techniques are used and the atmospheric parameter of interest is not measured directly. Instead, the radiation that is emitted, absorbed, or reflected by the atmospheric quan- tity of interest is measured. The remote sensing measurements can be either passive or active. Active remote sensing involves transmitting a source of electromagnetic radiation and receiv- ing the backscattered signal. Passive remote sensing is similar to the active method; however, a natural electromagnetic radiation source (such as the Sun or the ) is used. Ozonesondes are good examples of an in-situ measurement technique. Ozonesondes are attached to large weather balloons and measure the vertical ozone profile from the surface up to an altitude of approximately 35 km. These measurements have very high vertical resolutions (sample res- olution of 30 m). Furthermore, They can function under severe weather conditions and in all climate regions. Ozonesondes are providing high-quality data; however, they cannot reach altitudes higher than 35 km. Usually, ozonesondes data are archived either under the World Ozone and Ultraviolet radiation Data Center (WOUDC) network or the Southern Hemisphere ADditional OZonesondes (SHADOZ) network. The Electrochemical Concentration Cell is the most common type of ozonesonde and has been widely used at different locations (Kley et al., 1996; Schulz et al., 2001; Vomel¨ and Diaz, 2010). A chemical reaction between potassium iodide (KI) and ozone (O3) produces iodine molecules (I2). The concentration of I2 molecules is proportional to the ozone concentration. This chemical reaction between potassium iodide and ozone is the basis of ozonesonde measurements and is shown below:

2 KI + O3 + H2O −−−→ I2 + O2 + 2 KOH (R 44)

The iodine molecules will generate a current within the ozonesode cell. The current is directly proportional to the partial pressure of ozone in the sampled air:

PO3 = cTt100γ(I − Ib) (1.1)

−4 where PO3 is the ozone partial pressure, c = 4.309 ∗ 10 , T is the temperature of the sampled air, t100 is the time which is needed to pump 100 ml of air to the cell, γ is the efficiency of the pump, I is the produced current and Ib is the background current produced when there is no ozone. A detailed discussion on the topic can be found in Komhyr (1986) and Johnson et al. (2002). 1.6. Measurements of Atmospheric Ozone 13

The Dobson spectrometer is an example of a ground-based remote sensing instrument which is used for ozone measurements. As the Dobson spectrometer is a passive remote sens- ing instrument, it uses natural sources of light (like direct sunlight or diffuse light from clear or cloudy skies) for its ozone measurements. The ozone absorbs light at selected bands of the electromagnetic spectrum; thus the amount of light which is transmitted to the ground de- pends on the abundance of the amount of ozone along the line-of-sight, which is converted into the overhead ozone column. In this method, the ratio of sunlight intensity at two wave- lengths is measured. The ozone weakly absorbs one of these wavelengths whereas the ozone mostly absorbs the other one. Thus the method is based on the differential absorption method. The instrument measures the total column ozone which is the total amount of ozone inside a vertical column extending from the ground. The Brewer spectrometer is similar to the Dob- son spectrometer; however, it uses five or six wavebands. The Dobson instrument employs a selection of eight different wavelengths at UV band (from 305.5 nm to 339.8 nm). Using the Umkehr principle, the vertical ozone concentration at ten different altitude layers is retrieved. The Umkehr layers are approximately 5 km thick. Although the standard Umkehr method re- trieves at ten layers, the retrieval only contains four independent pieces of information (Mateer, 1965; Mateer and DeLuisi, 1992). Thus, using the Dobson instrument, long-term stratospheric ozone (from 20 km to 40 km) measurements are produced. However, these profiles have coarse vertical resolutions (between 5 km to 10 km).

Satellites can carry passive remote sensing instruments with the ability to provide global coverage for ozone measurements. Based on different geometric viewing concept, the electro- magnetic radiation which is reflected or emitted from the Earth’s atmosphere can be measured in nadir, limb, or occulation modes. Nadir measurements can provide good horizontal cov- erage; however, their vertical resolution is poor. Limb sounders can measure vertical profiles with each measurement representing a relatively narrow layer of atmosphere, thus compare to nadir scans, the vertical resolution of limb measurements is higher. However, due to clouds, aerosols, and in the troposphere, their measurement sensitivity at lower altitudes is poor. Furthermore, using solar, lunar, or stellar occultation techniques, high signal-to-noise ra- tio (SNR) measurements can be made. The advantage of the latter method is that it can measure troposphere ozone as well. However, solar occultation measurements can only be performed 14 Chapter 1. Introduction as the sun rises or sets (relative to their orbits). More detail on satellite measurements can be found elsewhere (e.g. Emery and Camps (2017)).

Lidar (LIght Detection And Ranging) is a ground-based active remote sensing technique which is similar to radar but operates using . In lidar measurements, a laser beam is sent into the atmosphere, and is scatters in all directions. A portion of the light is backscattered toward the lidar. A telescope collects the backscattered photons and is detected by photomulti- plier tubes (PMTs). The received signal contains information about the atmosphere (Weitkamp, 2006). In this thesis, we employed a Differential Absorption Lidar (DIAL) system to retrieve ozone density profiles. Details about the DIAL system and the ozone retrievals are presented in the next section.

1.7 DIAL Measurements and Retrieval of Ozone Density

A DIAL system is based on transmitting two wavelengths simultaneously to the atmosphere. One of the emitted wavelengths is strongly absorbed by the constituent of interest (called the “on-line” wavelength) and the other is weakly absorbed (called the “off-line” wavelength) and used as the reference wavelength. A schematic diagram of a DIAL system is shown in Fig 1.2. Here, we briefly describe how vertical ozone density profiles can be retrieved using a DIAL system. Detailed description on the topic can be found in Schotland (1974); Godin-Beekmann et al. (2003); Godin et al. (1999).

In ozone studies, selecting a wavelength pair depends on the altitude range of measure- ments, and wavelengths in the UV spectrum are the most efficient. A pair of wavelengths with a strong UV absorption is needed to detect the small amount of ozone which resides in the troposphere. However, for stratospheric ozone measurements, choosing a laser that can reach to higher altitudes in the stratosphere is the main concern (Megie et al., 1985; Browell, 1989; Papayannis et al., 1990). The general retrieval method for both tropospheric and stratospheric DIAL measurements is the same, and is based on the lidar equation in which the measured backscattered photocounts, Nobs(z,λ), for a laser pulse at wavelength λ and at altitude z can be written as (Schotland, 1974; Fernald, 1984; Weitkamp, 2006): 1.7. DIAL Measurements and Retrieval of Ozone Density 15

Figure 1.2: In the DIAL technique two wavelengths (the “on-line” and the “off-line” wave- lengths) are simultaneously transmitted to the atmosphere. The back scattered signals are col- lected by a large mirror or by multiple small mirrors. The collected signal (via optical fibers) are sent to the PMTs.

P A N(z, λ) = η exp(τ (z, λ)) exp(τ (z, λ))O(z) laser β ∆t∆z + B(z) (1.2) system emitted returned ~c 4πz2 λlaser

Each of the quantities in the above equation is listed below:

ηsystem: the efficiency of the lidar system

τemitted(z, λ): the optical depth of the emitted photon through the atmosphere

τreturned(z, λ): the optical depth of the backscattered photon through the atmosphere O(z): the geometrical overlap function of the lidar

Plaser: the power of laser which is used in the lidar ~: the Planck constant c: the speed of light

λlaser: the laser wavelength β: the atmospheric backscattering coefficient 16 Chapter 1. Introduction

A 4πz2 : the effective area of the primary telescope

∆t: the temporal resolution of the lidar

∆z: the spatial resolution of the lidar

B(z): the background counts which can be function of a altitude.

In Rayleigh scattering (for which the backscattered signal has the same wavelength as the transmitted signal), the optical depths for the transmitted and returned wavelengths are the same. In , because the backscattered wavelength is a Raman-shifted wave- length of the transmitted signal, these two terms are different. Moreover, as the probability of multiple scattering of photons is low (even the probability of back-scattering is so low), in the above equation, we assumed that each photon only back-scattered once. To retrieve the ozone density, in the clean (aersol free) atmospheric conditions, the Rayleigh scattering technique is used. However, in the presence of aerosols, the Raman scattering tech- nique is added to the retrieval processes. The focus of this thesis is on retrieving ozone density for “clean” nights, thus we only explain the Rayleigh technique. In Eq.1.2, the atmospheric optical depth (considering that the molecule of interest is ozone) is given by:

Z z 0 0 0 X 0 0 τ(z, λ) = [σO3 (λ, T(z ))nO3 (z ) + α(λ, z ) + σe(λ)ne(z )]dz (1.3) z0 e

where z0 is the altitude of the lidar station, σO3 (λi) is the ozone absorption cross section at the specific altitude and wavelength which is dependent to the atmospheric temperature 0 T(z ), nO3 (z) is the ozone number density to be measured, α(λ, z) is the atmospheric extinc- tion coefficient which includes both Rayleigh and Mie scattering extinction coefficients, and P e σe(λ)ne(z) is the the extinction by other absorbers (like SO2 and NO2). In major volcanic

eruptions the abundance of SO2 gas in the stratosphere can significantly perturb the ozone re-

trievals. However, SO2 only stays in the stratosphere for 30 to 40 days (Heath et al., 1983).

In general, the amount of SO2 mixing ratio in the stratosphere is negligible. The differential −19 2 absorption cross section of NO2 in the specified spectrum is on the order of 3 × 10 cm , thus considering the effect of NO2 in the ozone retrievals is not essential, and the third term of Eq.1.3 is negligible (Godin-Beekmann et al., 2003). 1.7. DIAL Measurements and Retrieval of Ozone Density 17

For a specific wavelength, in Eq.1.2, the efficiency, the laser power, and the effective area of primary telescope are constant values within a lidar system, thus we can write:

A P C(λ) = η laser (1.4) system 4πz2 ~c λlaser

where C(λ) is called the lidar constant.

1.7.1 Ozone Density Retrievals

By substituting Eq.1.3 and 1.4 into Eq.1.2 for “on-line” and “off-line” channels we can write:

Z z ! 0 0 0 0 Nobs(z, λon) =C(λon)β(λon) exp 2 [σO3 (λon, T(z ))nO3 (z ) + α(λon, z )]dz + B(z, λon) z0 Z z ! 0 0 0 0 Nobs(z, λo f f ) =C(λo f f )β(λo f f ) exp 2 [σO3 (λo f f , T(z ))nO3 (z ) + α(λo f f , z )]dz + B(z, λo f f ). z0 (1.5)

Dividing the two equations and taking the natural logarithm of the result gives:

! ! ! N (z, λ ) − B(z, λ ) C(λ ) β(λ ) Z z Z z ln obs on on = ln on + ln on + 2 ∆δ n (z0)dz0 + 2 ∆αdz0. − o3 O3 Nobs(z, λo f f ) B(z, λo f f ) C(λo f f ) β(λo f f ) z0 z0 (1.6)

In Eq. 1.6 and Eq.1.10 we used the term ∆δo3 which is the difference of ozone cross sections for the two different wavelengths:

∆δo3 = σo3(λon) − σo3(λo f f ). (1.7)

Similarly:

∆α = α(λon) − α(λo f f ). (1.8)

As mentioned earlier, the atmospheric extinction term (α) includes both molecules (indicating Rayleigh scattering) and particles (indicating Mie scattering), and it can be written as:

α(λ, z) = αp + Σn(z)σR(λ) (1.9) 18 Chapter 1. Introduction

where, αp is the particulate extinction coefficient, σR(λ) is the molecular Rayleigh cross section at a given wavelength, and n(z) is the number density of molecules. The term β(λ) also includes both molecular and particulate backscattering coefficients. Taking the derivative of Eq.1.6 with respect to the altitude and rearranging the equation, ozone density profile can be retrieved:

! ! 1 d N(λo f f , z) − Bo f f (z) 1 d β(λon, z) 1 1 no3(z) = ln − ln − ∆αp(z) − ∆σR. 2∆δo3 dz N(λon, z) − Bon(z) 2∆δo3 dz β(λo f f , z) ∆δo3 ∆δo3 (1.10) As the lidar constants are not functions of altitude, the derivative of them with respect to height is zero, and they do not play any role in calculating the ozone density profiles. The first term of the above equation contains the backscattered photon counts and background counts at each wavelength. The second term shows the ratio of molecular and particulate backscattering co- efficients for “on-line” and “off-line” wavelengths. For “clean” nights (when the amount of in the atmosphere is insignificant) this term is negligible:

! ! 1 d β(λ , z) 1 d σ (λ ) ln on = ln R on = 0 (1.11) 2∆δo3 dz β(λo f f , z) 2∆δo3 dz σR(λo f f )

In the third term, the difference between the particulate extinction coefficients for the two wavelengths is shown. Similar to the second term, for “clean” nights this term is negligible. The fourth term is the difference between the Rayleigh cross sections at the two wavelengths (i.e, the difference in the molecular extinction coefficients). To calculate this term the At- mospheric density profile should be known; at altitudes below 30 km density profiles from nearby are used. As radiosondes can not reach into higher altitudes, an atmo- spheric model is normally used. These models have high uncertainties; however, the ozone density uncertainty resulting from the errors on the air density above 15 km is less than 1%. The contribution of air density error below 15 km is important and at some cases can be as high as 15% (Megie et al., 1985). In summary, for ozone retrievals in a “clean” night condition (which is the focus of this thesis), the most important term of Eq.1.10 is the first term which contains the lidar measurements at the two wavelengths. However, prior to using the measure- ments to retrieve the ozone density profile, some corrections should be applied. In the next section, these corrections are explained in detail. 1.8. Corrections Applied to the Raw Counts 19

1.8 Corrections Applied to the Raw Counts

1.8.1 DeadTime Correction

For many lidar systems, at count rates below about 1 MHz, the relation between the true counts and the observed signal is linear. However, for the higher counts, the detector’s response may not be linear. The nonlinearity becomes more significant as the count rate increases. Detector systems are generally limited in their useful dynamic range due typically to the detector (par- alyzable) or counting system (non-paralyzable). In paralyzable systems, if the time interval between two photon strikes is shorter than the time needed to process the first photon strike (dead time γ) the detector is unable (paralyzed) to observe the second photon. This relation between the observed and true photon counts in paralyzable detectors is then:

Nobs = Ntrue exp(−γNtrue) (1.12)

where Nobs is the observed counts by the detector, Ntrue is the true counts (Donovan et al., 1995). In non-paralyzable systems, the detector is not paralyzed but the counter is unable to record another photon in the time interval of γ after any recorded photon strike. The relation between the true and observed counts for the system shows is then:

Ntrue Nobs = (1.13) 1 + γNtrue

The lidars used in this thesis have non-paralyzable counting systems. To use measurements in these systems above about 1 MHz the dead time must be specified or retrieved. In the traditional method, the lidar measurements should be corrected for the effect of deadtime. If the value of the deadtime is not known, an empirical fit can be used to estimate this value.

1.8.2 Signal Induced Noise

It is well-known that for high intensity systems, the output of the PMT can show an excess of counts some time after the signal intensity is maximum, a “tail” which is called signal-induced 20 Chapter 1. Introduction noise (SIN) (Hunt and Poultney, 1975). In fact, SIN is the residual signal originating from high signal intensities at low altitudes. It adds up with the background signal and is visible at altitudes where the signal-to-noise ratio (SNR) is very small (Iikura et al., 1987). Using a mechanical chopper to block high intensity light from reaching the detector is the most practical way to avoid SIN. It is important to consider the noise component from the upper altitude of lidar signals. In many lidars, the background is a constant and the effect of SIN is not detected. However, when the SIN is present in the background, the uppermost part of the signal can be fitted by an exponential function:

B(z) = a exp(−bz) + c (1.14) where the fitting coefficients a, b, and c are empirically determined (Iikura et al., 1987). The SIN is more pronounced for the “on-line” wavelength because most of the laser power is used for the “on-line” wavelength (normally, the “on-line” wavelength has a power about 2 times of the laser power in the “off-line” wavelength). Therefore, for most nights the affect of SIN on the “off-line” wavelength is negligible, and a constant background is used.

1.8.3 Merging Process

Due to the high dynamic range of signals, usually in lidar measurements, two detecting chan- nels are used. One channel is used for high altitude measurements while the other is optimized for the lower altitude measurements. Before applying the retrieval algorithm, the two high- altitude and low-altitude channels are merged to produce one signal. To merge the two signals an optimized height should be determined where both signals have the same SNR, and they are linear with respect to each other. A major issue appears when merging analog and digital chan- nels is required. The most common practice is to digitize the analog signal, and then merge the two channels. The uncertainty which is introduced in the conversion process is yet to be math- ematically determined. Moreover, for many lidars, the digitized analog signal does not follow the Poisson distribution, thus determining the gluing uncertainty becomes more difficult. A detailed description of the merging process can be found in Steinbercht (1994). 1.9. Optimal Estimation Method 21

1.8.4 Digital Filters

In the DIAL technique, the rapid decrease of the SNR is another difficulty. Low-pass filters are used to reduce the noise of the signals. The final vertical resolution ∆z f varies by the order of filter (number of point used to make the filter) and is calculated as:

∆z f = νc∆zi (1.15)

where νc is the cutoff frequency of the low-pass filter, and ∆zi is the initial vertical resolution of the measurements (Godin et al., 1999; Leblanc et al., 2016). In the lower stratosphere, perturbations in the ozone profiles are well detected; however, depending on the order of filter, the perturbation can be largely attenuated and cause negative or positive biases. For higher altitudes, because of the lower SNR, the vertical resolution is decreased. Different numerical filters have been tested to optimize the ozone retrievals. In all these techniques, to overcome the SNR decrease, the number of coefficients in filters are increased with altitude (Godin et al., 1999). In summary, for traditional DIAL analysis some corrections should be applied to raw count measurements, after which Eq.1.10 can be applied to calculate the ozone density profile. An alternative approach is to apply the Optimal Estimation Method (OEM). In this method, count correction, gluing of profiles, and pre- or post-filtering are not needed. The raw measurements from all the available channels are used as an input vector and one ozone profile is retrieved as the output. In the next section more detail on OEM is provided.

1.9 Optimal Estimation Method

Inverse modeling is a process of the transformation from data to model parameters. OEM is a matrix inverse method based on Bayes' theorem. Let x = (x1, x2, ..., xn) be the vector state of the atmosphere and y = (y1, y2, ..., yn) be the corresponding vector measurement. The relationship between x and y can be shown with a forward model:

y = F(x, b) +  (1.16) 22 Chapter 1. Introduction where b is the model parameter vector and  is the noise in the measurements. In the absence of error from the inversion of the forward model, the exact value of x can be retrieved. However, all real measurements contain experimental errors. Therefore, any practical retrieval method should carry the measurement's uncertainty and the resulting uncertainty of the retrieved quan- tity. Bayesian statistics provides a useful way to look at this problem.

In Bayes’ theorem, an a priori state xa and its assigned probability reflecting the certainty of this state is provided. The goal is to calculate the most likely state vector x which is consistent with the a priori knowledge. Formally we can write:

P(y|x)P(x) P(x|y) = . (1.17) P(y)

In the above equation:

• P(x) is the a priori probability density function (PDF) of the state vector x.

• P(y|x) is the PDF of the measurement vector y (given the true values of b).

• P(x|y) is the PDF of x (given the measurement vector y), and is called the a posteriori PDF for the state vector.

We use matrix presentation in which the uncertainties associated with the a priori state and

measurements are shown as covariance matrices Sa and Sy where their diagonal elements are the variances of the individual elements of the a priori state and measurements noise. Here, for simplicity, a linear problem in which all of the PDFs follow Gaussian statistics is assumed. In most altitudes we have more than 15 photons. Hence, according to the central limit theorem, assuming a Gaussian distribution is a valid assumption. The Gaussian distribution in vector space can be written as:

! 1 1 | −1 P(y) = n 1 exp − (y − y¯) Sy (y − y¯) . (1.18) (2π) 2 |S y| 2 2

∂F Assuming that the forward model is linear, we define the Jacobian matrix as Kx = ∂x which indicates the sensitivity of the forward model to the state variables x. Therefore, the PDF of 1.9. Optimal Estimation Method 23 the measurement vector y and the state vector x, after some rearrangements, is written as:

| −1 −2 ln P(y|x) = (y − y¯) Sy (y − y¯) + c1 (1.19) | −1 −2 ln P(x) = (x − xa) Sa (x − xa) + c2

where c1 and c2 are constants. The posteriori PDF is:

T −1 T −1 −2 ln P(x|y) = (y − Kx) Sy (y − Kx) + (x −bx) Sa (x −bx) (1.20) where −2 ln P(x|y) is called the cost function. The optimal or maximum posteriori (MAP) solution for x is shown as the the maximum of P(x|y) which is the solution to 5x(−2 ln P(x|y)) =

0 where 5x is the gradient operator in the state vector. In Eq.1.20, the first term is defining the difference between the true value (measurements) and the predicted value (the forward model) weighted by the measurement noise (error), thus this term is in fact the well-known Least Square Method which is widely used in regression problems. The second term of the equation defines the difference between an a priori value and the state vector weighted by a priori uncertainty. Depending on how much we trust the a priori profile (how large the a priori uncertainty is) the second term can play an important role. Typically, the cost is normalized to the number of measurements, and a cost of around 1 indicates a good retrieval.

The solution to 5x(−2 ln P(x|y)) = 0 (the MAP solution) is:

| −1 −1 | −1 −1 bx = xa + (K Sy K + Sa )K Sa (y − Kxa) (1.21) or equivalently:

bx = xa + G(y − Kxa) (1.22)

∂bx where G = ∂y is the gain matrix and represents the sensitivity of the retrieval to the observa- tions. Moreover, the averaging kernel matrix is defined as the sensitivity of the optimal solution to the true state x: ∂x A = b = G K (1.23) ∂x y x

In practice, most forward models are nonlinear and the Eq.1.20 should be solved numeri- 24 Chapter 1. Introduction cally. The Newton and Gauss-Newton methods are normally used when the problem is not too non-linear, and the Levenberg-Marquart iteration is used for forward models that have higher degree of nonlinearity. Here, we use the latter method. The optimized solution for state vector x is found when the below iteration converges:

−1 T −1 T −1 −1  xi+1 = xi + [(1 + γi)Sa + Ki SyKi] [Ki Sy (y − F(xi, b)] − Sa (xi − xa) . (1.24)

Here, γi is a damping factor, for γi −→ 0 the iteration becomes similar to the Gauss-Newton method, and for γi −→ inf the iteration tends to steepest descent. Detailed discussion on the choice of γi can be found elsewhere (Marquardt, 1963). Finally, the uncertainty budget can be calculated as:

Stotal = Sm + Ss + SF. (1.25)

In the latter equation, the retrieval covariance due to the measurement noise is S m:

T Sm = GySyGy . (1.26)

Rather than being the estimate of the true state, the retrieval is an optimal smoothed estimate (smoothed by the averaging kernel). The retrieval error resulting from smoothing should be calculated. The smoothing covariance Ss is:

T Ss = (A − In)Se(A − In) (1.27)

where, In is the unit matrix and Se is the covariance of the real ensemble of states. The error in the retrievals due to the forward model parameter uncertainties S F is defined as:

T T SF = GyKbSbKb Gy . (1.28)

dF In the above equation, Kb = db represents the sensitivity of forward model to the b parameter,

and Sb is the error on the assumed b parameter. A summary of the OEM procedure is shown in Fig.1.3. In Chapters 2 and 3, we describe how the OEM can be implemented to retrieve 1.10. Machine Learning Applications to Lidar Measurements 25 stratospheric and tropospheric ozone profiles, and we compare our results with the traditional analysis.

Figure 1.3: The flowchart is the summery of the steps we take to calculate the retrieval and its associated uncertainties.

1.10 Machine Learning Applications to Lidar Measurements

In a lidar system, the back scattered signals are received and recorded as level 0 measurement scans. Later on, to improve the SNR, these scans are co-added in time to produce one single profile representing a period of measurement. Prior to co-adding, each individual scan should be checked to make sure if it has a good quality. Therefore, “bad scans” will be removed from the “good scans”. In this process, scans with low laser signal, high background counts, or unusual shape are flagged as “bad scans”. Scans with traces of clouds or aerosols are also separated from the “clean and clear” scans. These scans are not bad scans, but they may require different processing algorithms. Scans are often classified manually as good or bad, and clean or not-clean. This method is time consuming and to some extent the classification is subject to the judgment of the observer. Some lidar groups also use simple automated routines in which a pre-defined thresh-hold for 26 Chapter 1. Introduction the SNR at a fixed altitude is used to classify good and bad scans. This method does not have high accuracy and bad scans may pass the thresh-hold test and incorrectly be flagged as good. Using state-of-the-art machine learning (ML) techniques, we have developed an automated classifier. We are classifying the level 0 lidar measurements with high accuracy. ML has recently been used to distinguish between aerosols and clouds for the -Aerosol Lidar and Infrared Pathfinder Satellite Observations (CALIPSO) level 2 measurements (Zeng et al., 2018). Furthermore, Nicolae et al. (2018) used a Neural Network algorithm to estimate the most probable aerosol types in a set of data obtained from European Aerosol Research Lidar Network (EARLINET). We present our result for the Purple Crow Lidar (PCL) system as well as for the Meteoswiss Raman Lidar for Meteorological Observations (RALMO) system. Here, we briefly describe the ML technique.

1.10.1 Machine Learning Methods

The ML is widely used for making predictions and to recognize patterns. The ML is catego- rized into supervised and unsupervised methods. In the supervised approach, the aim of ML is to learn a function which can map observations (x) to correct dependent (output) values (y). In supervised learning the dataset is divided into the training set and the test set. In the traning phase, the training set (which contains x values and their corresponding y values) is used to learn the mapping function. To validate the result, the test set is used (which only has x values) to predict the output. A good supervised algorithm will produce high accuracy scores in both training and test phases. Unsupervised learning is a data driven method in which the main goal is to find similarities between data points and to cluster the data accordingly. Thus output values (y) are not needed and the algorithm is trained only by using the observations (x). It is worth noting that in inverse modelling methods (e.g., OEM), y is considered as the observable and x is the quantity of interest. However, in ML notation, x is given and y is predicted. The ML can be used for either classification or regression. In the classification methods, targets are discrete values or categories. For example, our classifier method in which the scans are divided into good and bad scans uses a classification ML model. In regression, the target is a continuous quantity. For example, a regression model can be used to retrieve the ozone 1.10. Machine Learning Applications to Lidar Measurements 27 density profiles. In Chapter 4 of this thesis, we use both supervised and unsupervised methods to classify our lidar measurements. We use different ML algorithms in our work and those which provided us with high accuracy scores are selected and discussed in detail. 28 Chapter 1. Introduction

Bibliography

Ahrens, C. D.: Meteorology Today: an introduction to weather, climate, and the environment, Meteorologie, 9, 74–75, 1998.

Andrews, D. G., Holton, J. R., and Leovy, C. B.: Middle atmosphere dynamics, 40, Academic press, 1987.

Ball, W. T., Alsing, J., Mortlock, D. J., Staehelin, J., Haigh, J. D., Peter, T., Tummon, F., Stubi,¨ R., Stenke, A., Anderson, J., Bourassa, A., Davis, S. M., Degenstein, D., Frith, S., Froidevaux, L., Roth, C., Sofieva, V., Wang, R., Wild, J., Yu, P., Ziemke, J. R., and Rozanov, E. V.: Evidence for a continuous decline in lower stratospheric ozone offsetting ozone layer recovery, acp, 18, 1379–1394, 2018.

Bates, D. R. and Nicolet, M.: Atmospheric hydrogen, Publications of the Astronomical Society of the Pacific, 62, 106–110, 1950.

Brasseur, G. P. and Solomon, S.: Aeronomy of the middle atmosphere: chemistry and physics of the stratosphere and mesosphere, vol. 32, 2006.

Browell, E. V.: Differential absorption lidar sensing of ozone, IEEE, 77, 419–432, 1989.

Chapman, S.: XXXV. On ozone and atomic oxygen in the upper atmosphere, The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science, 10, 369–383, 1930.

Cohen, Y., Petetin, H., Thouret, V., Marecal,´ V., Josse, B., Clark, H., Sauvage, B., Fontaine, A., Athier, G., Blot, R., et al.: Climatology and long-term evolution of ozone and carbon monoxide in the upper troposphere–lower stratosphere (UTLS) at northern midlatitudes, as seen by IAGOS from 1995 to 2013, acp, 18, 5415–5453, 2018.

Crutzen, P. J.: The influence of nitrogen oxides on the atmospheric ozone content, Quarterly Journal of the Royal Meteorological Society, 96, 320–325, 1970.

Danielsen, E. F.: Stratospheric-tropospheric exchange based on radioactivity, ozone and poten- tial vorticity, 1968. BIBLIOGRAPHY 29

Donovan, D. P., Bird, J. C., Whiteway, J. A., Duck, T. J., Pal, S. R., and Carswell, A. I.: Lidar observations of stratospheric ozone and aerosol above the Canadian high arctic during the 199495 winter, Geophysical Research Letters, 22, 34893492, 1995.

Emery, W. and Camps, A.: Introduction to Satellite Remote Sensing: Atmosphere, Ocean, Land and Cryosphere Applications, Elsevier, 2017.

Farman, J. C., Gardiner, B. G., and Shanklin, J. D.: Large losses of total ozone in Antarctica reveal seasonal ClOx/NOx interaction, Nature, 315, 207, 1985.

Fernald, F. G.: Analysis of atmospheric lidar observations: some comments, Appl. Opt., 23, 652–653, 1984.

Finlayson-Pitts, B. J., Jr., P., and N., J.: Chemistry of the upper and lower atmosphere: theory, experiments, and applications, Elsevier, 1999.

Gebhardt, C., Rozanov, A., Hommel, R., Weber, M., Bovensmann, H., Burrows, J. P., Degen- stein, D., Froidevaux, L., and Thompson, A. M.: Stratospheric ozone trends and variability as seen by SCIAMACHY from 2002 to 2012, acp, 14, 831–846, 2014.

Godin, S., Carswell, A. I., Donovan, D. P., Claude, H., Steinbrecht, W., McDermid, I. S., McGee, T. J., Gross, M. R., Nakane, H., Daan, Swart, P. J., Bergwerff, B. B., Uchino, O., von der Gathen, P., and Neuber, R.: Ozone differential absorption lidar algorithm intercom- parison, Appl. Opt., 38, 6225–6236, 1999.

Godin-Beekmann, S., Porteneuve, J., and Garnier, A.: Systematic DIAL lidar monitoring of the stratospheric ozone vertical distribution at Observatoire de Haute-Provence (43.92[de- gree]N, 5.71[degree]E), J. Environ. Monit., 5, 57–67, 2003.

Harris, N. R. P., Hassler, B., Tummon, F., Bodeker, G. E., Hubert, D., Petropavlovskikh, I., Steinbrecht, W., Anderson, J., Bhartia, P. K., Boone, C. D., Bourassa, A., Davis, S. M., Degenstein, D., Delcloo, A., Frith, S. M., Froidevaux, L., Godin-Beekmann, S., Jones, N., Kurylo, M. J., Kyrol¨ a,¨ E., Laine, M., Leblanc, S. T., Lambert, J.-C., Liley, B., Mahieu, E., Maycock, A., de Maziere,` M., Parrish, A., Querel, R., Rosenlof, K. H., Roth, C., Sioris, C., 30 Chapter 1. Introduction

Staehelin, J., Stolarski, R. S., Stubi,¨ R., Tamminen, J., Vigouroux, C., Walker, K. A., Wang, H. J., Wild, J., and Zawodny, J. M.: Past changes in the vertical distribution of ozone Part 3: Analysis and interpretation of trends, acp, 15, 9965–9982, 2015.

Harrison, R. M.: Understanding our environment: an introduction to environmental chemistry and pollution, 2007.

Heath, D. F., Schlesinger, B. M., and Park, H.: Spectral change in the ultraviolet absorption and scattering properties of the atmosphere associated with the eruption of El Chichon:´ Strato- spheric SO2 budget and decay, Eos Trans. AGU, 64, 197, 1983.

Hedin, A. E.: Extension of the MSIS thermosphere model into the middle and lower atmo- sphere, J. Geophys. Res.: Space Physics, 96, 1159–1172, 1991.

Holton, J. R., Haynes, P. H., McIntyre, M. E., Douglass, A. R., Rood, R. B., and Pfister, L.: Stratosphere-troposphere exchange, Rev. Geophys., 33, 403–439, 1995.

Hunt, W. H. and Poultney, S. K.: Testing the linearity of response of gated photomultipliers in wide dynamic range laser radar systems, IEEE Trans. Nucl. Sci, 22, 116–120, 1975.

Iikura, Y., Sugimoto, N., Sasano, Y., and Shimzu, H.: Improvement on lidar data processing for stratospheric aerosol measurements, Appl. Opt., 26, 5299–5306, 1987.

IPCC2007: Climate Change 2007:Th Physical ScienceSummary for Policy makers. Contribu- tion of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change,17.

Jacob, D. J., Heikes, E., Fan, S. M., Logan, J. A., Mauzerall, D. L., Bradshaw, J. D., Singh, H. B., Gregory, G. L., Talbot, R. W., Blake, D. R., et al.: Origin of ozone and NOx in the tropical troposphere: A photochemical analysis of aircraft observations over the South Atlantic basin, J. Geophys. Res.: Atmospheres, 101, 24 235–24 250, 1996.

Johnson, B. J., Oltmans, S. J., Vomel,¨ H., Smit, H. G., Deshler, T., and Kroger,¨ C.: Electro- chemical concentration cell (ECC) ozonesonde pump efficiency measurements and tests on BIBLIOGRAPHY 31

the sensitivity to ozone of buffered and unbuffered ECC sensor cathode solutions, Journal of Geophysical Research: Atmospheres, 107, ACH–8, 2002.

Johnston, H. S.: Atmospheric ozone, Annual review of physical chemistry, 43, 1–31, 1992.

Junge, C. E.: Global ozone budget and exchange between stratosphere and troposphere, Tellus, 14, 363–377, 1962.

Kley, D., Crutzen, P. J., Smit, H. G. J., Vomel,¨ H., Oltmans, S. J., Grassl, H., and Ramanathan, V.: Observations of near-zero ozone concentrations over the convective Pacific: Effects on air chemistry, Science, 274, 230–233, 1996.

Komhyr, W.: Operations handbook-ozone measurements to 40-km altitude with Model 4A Electrochemical Concentration Cell (ECC) ozonesonaes (used with 1680-MHz radioson- des), Unknown, 1986.

Leblanc, T., Sica, R. J., van Gijsel, J. A. E., Godin-Beekmann, S., Haefele, A., Trickl, T., Payen, G., and Gabarrot, G.: Proposed standardized definitions for vertical resolution and uncertainty in the NDACC lidar ozone and temperature algorithms Part 1: Vertical resolu- tion, amt, 9, 4029–4049, 2016.

Manney, G. L., Santee, M. L., Rex, M., Livesey, N. J., Pitts, M. C., Veefkind, P., Nash, E. R., Wohltmann, I., Lehmann, R., Froidevaux, L., et al.: Unprecedented Arctic ozone loss in 2011, Nature, 478, 469, 2011.

Marquardt, D. W.: An algorithm for least-squares estimation of nonlinear parameters, Journal of the society for Industrial and Applied Mathematics, 11, 431–441, 1963.

Mateer, C. L.: On the information content of Umkehr observations, J. Atmospheric Sci., 22, 370–381, 1965.

Mateer, C. L. and DeLuisi, J. J.: A new Umkehr inversion algorithm, J. Atmospheric Sol.-Terr. Phys., 54, 537–556, 1992.

McElroy, M. B., Salawitch, R. J., Wofsy, S. C., and Logan, J. A.: Reductions of Antarctic ozone due to synergistic interactions of chlorine and bromine, Nature, 321, 759, 1986. 32 Chapter 1. Introduction

Megie, G. J., Ancellet, G., and Pelon, J.: Lidar measurements of ozone vertical profiles, Appl. Opt, 24, 3454–3463, 1985.

Molina, M. J. and Rowland, F. S.: Stratospheric sink for chlorofluoromethanes: chlorine atom- catalysed destruction of ozone, Nature, 249, 810, 1974.

Nicolae, D., Vasilescu, J., Talianu, C., Binietoglou, I., Nicolae, V., Andrei, S., and Antonescu, B.: A neural network aerosol-typing algorithm based on lidar data, Atmospheric Chemistry and Physics, 18, 14 511–14 537, 2018.

Papayannis, A., Ancellet, G., Pelon, J., and Megie, G.: Multiwavelength lidar for ozone mea- surements in the troposphere and the lower stratosphere, Appl. Opt, 29, 467–476, 1990.

Ramaswamy, V., Boucher, O., Haigh, J., Hauglustine, D., Haywood, J., Myhre, G., Nakajima, T., Shi, G. Y., and Solomon, S.: Radiative forcing of climate, Climate change, 349, 2001.

Sauvage, F., Laffont, L., Tarascon, J. M., and Baudrin, E.: Study of the insertion/deinsertion mechanism of sodium into Na0. 44MnO2, Inorganic chemistry, 46, 3289–3294, 2007.

Schotland, R.: Errors in the lidar measurement of atmospheric gases by differential absorption, Journal of Applied Meteorology (1962-1982), pp. 71–77, 1974.

Schulz, A., Rex, M., Harris, N. R. P., Braathen, G. O., Reimer, E., Alfier, R., Kilbane-Dawe, I., Eckermann, S., Allaart, M., Alpers, M., et al.: Arctic ozone loss in threshold condi- tions: Match observations in 1997/1998 and 1998/1999, J. Geophys. Res.: Atmospheres, 106, 7495–7503, 2001.

Solomon, S., Ivy, D. J., Kinnison, D., Mills, M. J., Neely, R. R., and Schmidt, A.: Emergence of healing in the Antarctic ozone layer, Science, p. aae0061, 2016.

Steinbercht, W.: Lidar measurements of ozone, aerosol and temperature in the stratosphere, Ph.D. thesis, York University, 1994.

Stolarski, R. S. and Cicerone, R. J.: Stratospheric chlorine: a possible sink for ozone, Can. J. Chem., 52, 1610–1615, 1974. BIBLIOGRAPHY 33

Tung, K., Ko, M. K., Rodriguez, J. M., and Sze, N.: Are Antarctic ozone variations a manifes- tation of dynamics or chemistry?, Nature, 322, 811, 1986.

Vomel,¨ H. and Diaz, K.: Ozone sonde cell current measurements and implications for observa- tions of near-zero ozone concentrations in the tropical upper troposphere, amt, 3, 495–505, 2010.

Weitkamp, C.: Lidar: range-resolved optical remote sensing of the atmosphere, vol. 102, 2006.

WMO: Scientic Assessment of Ozone Depletion: 1988 Global Ozone Research and Monitoring Project Report 44, World Meteorological Organization, Geneva, 1988.

WMO: Scientic Assessment of Ozone Depletion: 1992 Global Ozone Research and Monitoring Project Report 40, World Meteorological Organization, Geneva, 1992.

WMO: Scientic Assessment of Ozone Depletion: 1998 Global Ozone Research and Monitoring Project Report 44, World Meteorological Organization, Geneva, 1999.

WMO: Scientic Assessment of Ozone Depletion: 2010 Global Ozone Research and Monitoring Project Report 52, World Meteorological Organization, Geneva, 2011.

WMO: Scientic Assessment of Ozone Depletion: 2014 Global Ozone Research and Monitoring Project Report, World Meteorological Organization, Geneva, 2014.

Zeng, S., Vaughan, M., Liu, Z., Trepte, C., Kar, J., Omar, A., Winker, D., Lucker, P., Hu, Y., Getzewich, B., and Avery, M.: Application of High-Dimensional Fuzzy K-means Clus- ter Analysis to CALIOP/CALIPSO Version 4.1 Cloud-Aerosol Discrimination, amt Discus- sions, 2018, 1–40, 2018. Chapter 2

Implementing the OEM to Retrieve Stratospheric Ozone Density

2.1 Overview

This chapter provides a detailed description of the first principle Optimal Estimation Method (OEM) which is applied to ozone retrieval analysis using Differential Absorption Lidar (DIAL) measurements. The air density, detector dead times, background coefficients, and lidar con- stants are simultaneously retrieved along with ozone density profiles. Using an averaging ker- nel, the OEM provides the vertical resolution of the retrieval as a function of altitude. A maximum acceptable height at which the a priori has a small contribution to the retrieval is calculated for each profile as well. Moreover, a complete uncertainty budget including both systematic and statistical uncertainties is given for each individual retrieved profile. Long term stratospheric DIAL ozone measurements have been carried out at the Observatoire de Haute- Provence (OHP) since 1985. The OEM is applied to three nights of measurements at OHP during an intensive ozone campaign in July 2017 for which coincident lidar-ozonesonde mea- surements are available. The retrieved ozone density profiles are in good agreement with both traditional analysis and the ozonesonde measurements. For the three nights of measurements, below 15 km the difference between the OEM and the sonde profiles is less than 25%, at alti- tudes between 15 km to 25 km the difference is less than 10%, and the OEM can successfully capture many variations of ozone which are detected in the sonde profiles due to its ability to

34 2.2. Ozone Retrievals:Traditional versus the OEM 35 adjust its vertical resolution as the signal varies. Above 25 km the difference between the OEM and the sonde profiles does not exceed 20%.

2.2 Ozone Retrievals: Traditional versus the OEM

In the traditional method, as discussed in Section 1.7.1, for retrieving ozone density from DIAL measurements, the derivative of the ratio between the “on-line” and “off-line” signals is calculated. By rewriting Eq. 1.10, the ozone number density can be retrieved as follows:

! −1 d N(λon, z) − Bon(z) no3(z) = ln + δno3 (z) (2.1) 2∆δo3(z) dz N(λo f f , z) − Bo f f (z) where N(λon, z) and N(λo f f , z) are, respectively, the “on-line” and “off-line” signals at altitude z, Bon(z) and Bo f f (z) are the background signals, and ∆δo3 (z) is the differential absorption cross section between the two wavelengths. δno3 (z) is a correction term for the effect of differential Rayleigh and Mie scattering, and the differential absorption by other absorbers (this term is equivalent to the last three terms of Eq. 1.10). More details can be found in McDermid et al. (1990), citeB205880D, and Leblanc et al. (2016b).

In the traditional ozone retrieval algorithm, several corrections are applied to the raw (level 0) counts to produce corrected photocounts, as discussed in Section 1.8. For high count rates, the dead time of the counting system is determined and a non-linearity correction is applied. Depending on the configuration of the lidar, channels with different gains may be merged (”glued”) to produce a single ozone profile. Determining the optimized height to merge the channels is typically done empirically. In the DIAL technique, the rapid decrease of sensitivity to ozone in the upper stratosphere is another important consideration. Low-pass filters are used to reduce the noise of the signals. For an ideal low-pass filter, the transfer function of all frequencies between 0 and the cut-off frequency, νc is 1, and the transfer function from νc to 1 f is 0, where the reduced frequency ν is defined as and fN is the Nyquist frequency. The final fN vertical resolution of the signal, ∆z f , varies by the order of filter, which depends on the cutoff 36 Chapter 2. Implementing the OEM to Retrieve Stratospheric Ozone Density

frequency and the initial vertical resolution ∆zi:

∆z f = νc∆zi. (2.2)

A detailed discussion on the digital filtering and the vertical resolution can be found in Godin et al. (1999) and Leblanc et al. (2016a).

In the lower stratosphere, the perturbations in the ozone profiles are well detected; how- ever, depending on the number of points in the filter (order of filter), the perturbation can be largely attenuated and cause negative or positive biases. For higher altitudes, because of the lower SNR, the vertical resolution is decreased. Different numerical filters have been tested to optimize the ozone retrievals. In all these techniques, to overcome the SNR decrease, the number of coefficients in the filters is increased with altitude (Godin et al., 1999).

2.2.1 Applying the optimal estimation method to ozone retrievals

The OEM is an inverse method in which the Bayesian theorem is used to find the probability distribution function (PDF) of the state of interest. Let x = (x1, x2, ..., xn) be the state vector, and y = (y1, y2, ..., yn) be the vector of the measurements. The relation between the measurements and the state vector is: y = F(x, b) +  (2.3) where F(x, b) is called the forward model. The forward model describes our understanding of the physics of the measurements as well as the instrument’s characteristics. Here, b is the model parameter vector which contains additional parameters needed in the forward model, and the noise in the measurements is the vector . In lidar measurements, the photon counts follow a Poisson distribution. However, for a count rate greater than 10 to 20, the PDF of the corresponding error tends toward a Gaussian distribution. Therefore, using the Bayesian approach and assuming a Gaussian PDF for all quantities, for a given measurement y, the most likely state of x is found by minimizing the following cost function:

T −1 T −1 J(x) = [y − F(bx, b)] Sy [y − F(bx, b)] + [bx − xa] Sa [bx − xa] (2.4) 2.2. Ozone Retrievals:Traditional versus the OEM 37

where Sy is the covariance matrix of the measurements, xa is the a priori profile which is an

initial guess for the state vector, and Sa is the associated a priori covariance matrix. Typically, the cost is normalized to the number of measurements, and a cost of around 1 indicates a good retrieval.

As the forward model is nonlinear, the MarquardtLevenberg method is used to find the state vector. The optimized solution for the state vector x occurs when the following iteration converges:

−1 T T −1 T −1 −1  xi+1 = xi + [(1 + γi)Sa + Ki SyKi ] [Ki Sy (y − F(xi)] − Sa (xi − xa) (2.5)

dF here, K = dx is the Jacobian of the forward model, and γi is a damping factor for the iteration. A comprehensive description on the application of the Marquardt-Levenburg method to OEM can be found in (Rodgers, 2000).

2.2.2 Ozone DIAL Forward Model

Our first-principle OEM retrieval uses the lidar equation as the forward model and the raw counts are the measurements. The lidar equation for the true counts is:

C N (z, λ ) =( λon )β(z, λ )Γ (λ , z)Γ (λ , z) + B (z) true on z2 on O3 on atm on λon (2.6) Cλ N (z, λ ) =( o f f )β(z, λ )Γ (λ , z)Γ (λ , z) + B true o f f z2 o f f O3 o f f atm o f f λo f f (z)

where λon and λo f f represents the “on-line” and “off-line” channels, ΓO3 (λon,o f f , z) and Γatm(λon,o f f , z)

are respectively, the ozone and atmospheric transmissions in each wavelength, Cλon and Cλo f f

are the lidar constants, and Bλon(z) and Bλo f f (z) are the background counts. For the stratospheric ozone measurements, in the altitude region of retrieval, the overlap is complete, and thus, we have not included it in our forward model. Depending on the characteristics of the data acqui- sition system, the true counts are related to the observed counts by either Eq. 1.13 or 1.12. In multi-channel systems, our forward model calculates the “on-line” and “off-line” wavelengths 38 Chapter 2. Implementing the OEM to Retrieve Stratospheric Ozone Density

for both high altitude and low altitude channels. The transmissions are defined as:

− 2τO3,atm ΓO3,atm(λi, z) = e (2.7)

where the optical depth τO3,atm is previously defined in Eq. 1.3. Both atmospheric optical depth and atmospheric backscattering coefficients have contributions due to scattering from molecules and aerosols:

Z z τatm = τmol + τaer = [σRnair(z) + α(z)]dz (2.8) z0

βatm = βair(z) + βaer(z) (2.9) where βair(z) and βaer(z) are the corresponding air and aerosol backscattering coefficients. The “on-line” and “off-line” coefficients are related through the following equation:

λ o f f −a βaer(λo f f ) = βaer(λon) ) (2.10) λon where for aerosols the Ångstrom coefficient a equals approximately 1, and for molecular scat- tering the Ångstrom coefficient a equals 4. In this Chapter, we only considered the clean-night condition. Therefore, the aerosol contribution to the process is not included, but could be in the future.

Due to the presence of SIN in the “on-line” channel, the background is assumed to be a function of height in the form of Eq. 1.14, while due to a negligible presence of SIN in the “off-line” channel, a constant background is used. If necessary, it is possible and easy to assign any analytic function for the background in both channels. Therefore, if needed the background for the “off-line” channel can be assumed as a function of height as well. Using the above forward model, the ozone and air density profiles, the background coefficients, the dead time and the lidar constants for the 4 channels are simultaneously retrieved. Other parameters in the forward model are treated as model parameters. Hence, they are fixed but considered as a source of uncertainty contributing to the retrieved quantities (see Table. 2.1). 2.3. Implementing the optimal estimation method retrieval 39

2.3 Implementing the optimal estimation method retrieval

To find the optimize solution of Eq. 3.7, a priori profiles for ozone and air density, as well as a priori values for background counts, dead time, and lidar constants are needed. Furthermore, b model parameter values and the covariance matrix of the measurements, a priori profiles, and model parameters need to be calculated. A summary of steps needed to implement the OEM for our ozone retrievals is shown in Fig. 2.1. A detailed description of these steps is provided in this section.

Figure 2.1: To implement the OEM, the a priori profiles for ozone and air density, background counts, dead time values, and lidar constants are needed. Moreover, b parameters should be identified and proper values for them should be calculated. The covariance matrices for a priori profiles, measurements, and b parameters need to be calculated as well.

The a priori ozone profile used for all retrievals is from an OHP ozone climatology. The climatology contains monthly-averaged ozone profile using the last 30 years of OHP DIAL and SAGE II satellite overpass measurements. The standard deviation to the 2σ level for this 40 Chapter 2. Implementing the OEM to Retrieve Stratospheric Ozone Density

climatology is 50% below 25 km and 10% above 20 km altitude. Alternatively, we have used the U.S. Standard model (Krueger and Minzner, 1976) as a priori ozone profile, which yields similar results for our ozone retrievals. In the traditional method, the ratio of the “on-line” to “off-line” channels is calculated. Thus, there is no need to assume an air density profile to retrieve ozone. However, in the correction term (Eq. 2.1), the air density profile is needed and an atmospheric model or a measurement is used. In the OEM, we are retrieving the air density as a state vector, and the Mass Spectrometer Incoherent Scatter Radar (MSIS) air density profile is used as a priori profile. The MSIS profiles are generally in good agreement with the ozonesonde measurements of air density. An uncertainty of 15% is assigned to the a priori of air density. In the case of ozone and air density there is a vertical correlation between the elements of retrieval states. This corresponds to the off-diagonal elements of the a priori covariance matrix. Generally, it is difficult to quantify the vertical length of this correlation. We have used a correlation length (łs) of 1000 m for ozone at altitudes below 18 km and the correlation length of 1400 m at higher altitudes. The air density has a correlation length of 1400 m for all regions. A tent function is used to model the decay of correlation (?). For the “off-line” channel the mean of the counts above 80 km are taken as a priori back- grounds, and their variances divided by the number of bins in the selected altitude region is used as a priori uncertainties in the background counts. For the “on-line” channel, an expo- nential function in the form of Eqn. 1.14 is fitted to counts above 80 km. The coefficients of the function are the a priori values. Depending on how good the initial fit is, uncertainties are assigned to the a priori coefficients, but for most nights a 20% uncertainty is chosen. Using the forward model, the a priori lidar constants for both channels were estimated and an initial standard deviation of 10% for both channels is assigned. In a range in which photon counting measurements are linear (or non-linearity is correctable), Poisson statistics is applied. Thus, the measurement variances are the number of photons in each atmospheric layer located at altitude ∆z, and there is no correlation between different layers (the off-diagonal elements of the matrix are zero). The following quantities are calculated for the b parameters in the forward model. The Rayleigh extinction which is calculated using the Nicolet formula (Nicolet, 1984), and the 2.4. Application of the OEM to measurements from the OHP stratospheric ozone lidar 41 temperature-dependent ozone absorption coefficients, as suggested by (Orphal et al., 2016), are calculated based on the BrionDaumontMalicet (BDM) database (Malicet et al., 1995). Un- certainties of 0.3% and 2% (Leblanc et al., 2016a) are respectively assigned to the Rayleigh and ozone cross sections. The ozone absorption cross-section is a function of temperature. The BDM database provides values for 5 different temperatures; in order to find the ozone cross section for the whole region where ozone is retrieved, the temperature is interpolated. For the interpolation, the sonde temperature profiles are used at lower altitudes (up to the altitude at which sonde measurements are available) ,and the MSIS temperature profiles are used for higher altitudes. Thus, the effect of temperature uncertainty on the ozone cross section and the final retrievals needs to be calculated as well. An uncertainty of 19 K is assigned to sonde measurements of temperature, and an uncertainty of 35 K is used for the MSIS profiles. The covariance matrix of the b parameters will be used later to calculate the systematic uncertainty of the retrieved quantities. Values and associated uncertainties of the a priori profiles for the parameters which we are retrieving, as well as the forward model parameters which are considered as fixed parameters, (and thus, are not being retrieved) are summarized in Table. 2.1. As mentioned earlier, we are testing our model on a reasonably clear night condition from a high altitude site, therefore, we are assuming that the effects of aerosols are negligible. After calculating Sy, Sa, Sb, xa, and b values, we used the Qpack software for our OEM retrieval. Qpack is a free Matlab package designed for forward and inverse modelling (Eriksson et al., 2005).

2.4 Application of the OEM to measurements from the OHP stratospheric ozone lidar

OHP is located in the south of France at (44◦N, 6◦E, 650 m ASL). Long term stratospheric ozone DIAL measurements have been performed since 1985. In addition, the OHP lidar is part of the international Network for the Detection of Atmospheric Composition Change (NDACC). In the OHP DIAL system, the “on-line” wavelength is provided by a XeCl excimer laser emit- ting at 308 nm with an emission energy of 200 mJ and a repetition rate of 100 Hz. The “off-line” 42 Chapter 2. Implementing the OEM to Retrieve Stratospheric Ozone Density

Parameter Value Standard Deviation Measurements measured Poisson statistics Retrieved a priori values Ozone density OHP climatology 50% to 10% Air density MSIS 15% Dead time empirical fitting 20% Background (“off-line”) mean above 80 km standard deviation above 80 km Coefficients of SIN (“on- empirical fitting above 80 km 20% line”) Lidar constants estimate from FM 20% Forward model parameters Rayleigh-scatter cross section Nicolet 1984 0.3% Ozone absorption cross sec- BDM 1986 2% tion Temperature profile sonde measurements 19 K Temperature profile MSIS 35 K

Table 2.1: Values and associated uncertainties for the retrieved and forward model parameters. wavelength is generated from the third harmonic (355 nm) of a Continuum Nd:Yag laser, with an output energy of 40 mJ and the repetition rate of 50 Hz. In the receiving end of the DIAL system, four similar F/3 mirrors of 0.53 m diameter collect the backscattered signals. The alti- tude steps of measurements is 150 m. The collected signal is separated to the Rayleigh signals at the transmitted wavelengths (308 nm and 353 nm), and the corresponding 1st Stokes wave- lengths in the nitrogen Raman spectrum (332.8 nm and 386.7 nm). Furthermore, to handle the high dynamic range of lidar signals in the whole altitude range, the Rayleigh signals are sep- arated to the high and low gain channels. More details on the instrumentation can be found elsewhere (Godin-Beekmann et al., 2003). The optical fibres transmit the receiving signals to the optical analysis device. The signals are detected by bialkali PMTs (Hamamatsu R2693P). The photon counting systems become nonlinear in the lowermost stratosphere. To correct for the saturation effect the following equation is used:

Nc = 1 + ((1 − x)Nr − 1) exp(−xNr) (2.11) where Nc is the observable counts, Nr is the true counts, and x is an adjustment parameter 2.4. Application of the OEM to measurements from the OHP stratospheric ozone lidar 43 which equals the inverse of the maximum observed counts which is the definition of the dead time (Pelon and Megie,´ 1982). To correct for the saturation, using Eq.2.11, the parameter x is adjusted for each wavelength in order to get a best agreement between the slopes of high and low altitude signals. The altitude at which the two profiles are combined can vary from night to night (Godin-Beekmann et al., 2003). For the two wavelengths and two different altitude channels, the dead time can differ. Therefore, we are retrieving the dead times for each altitude and at each wavelength. A dead time value which corresponds to the x parameter of each channel at each night is used as our a priori, and an uncertainty of ±20% is assigned to it. Using the OEM, we retrieve the ozone density and air density profiles, as well as the dead time values for the four channels, the background counts for the “off-line” channel, and the SIN coefficients (three values) for the “on-line” channel. In total, we retrieved eight quantities along with the ozone density and air density profiles. The degree of freedom for our measurements, which is the trace of the averaging kernel, is ≈ 78. Below we present the ozone retrieval for 26 July 2017 in detail. In order to show that the OEM is a robust method, the results for the nights of 14, and 20 of July are presented as well. In all these nights, ozonesonde balloons were coincidentally launched, thus the OEM is validated against both the traditional method and the sonde measurements.

2.4.1 Applying the OEM to OHP measurements on 26 July 2017

Figure. 2.2 shows the averaged counts over 4 hours of measurements for two different chan- nels at “on-line” and “off-line” wavelengths on the night of 26 July 2017. The coincident ozonesonde is launched within one hour after the start of the measurement, and takes approx- imately 2 hours to reach 30 km. For each retrieval, the averaging kernel matrix is calculated. The averaging kernel is a diagnostic variable which describes how the retrieval sees changes in the real atmosphere. Therefore, it contains information on the sensitivity (area of the aver- aging kernel function) and on the smoothing (shape of the averaging kernel function) of the retrievals. Ideally the averaging kernel is a unity matrix preserving any change in the retrieved quantity from the a priori state. The area is defined as the vector product Au where u is a unit vector. When the retrieval comes solely from the measurements then the area equals 1, and at 44 Chapter 2. Implementing the OEM to Retrieve Stratospheric Ozone Density

altitudes where the a priori profile is contributing to the retrievals the area decreases, where an area equal to 0 would mean nothing is being retrieved.

Figure 2.2: Average count rates for 5 hours of measurements on 26 July 2017. Left panel:“on- line” wavelength (blue curve, low altitude; red curve, high altitude). Right panel: “off-line” wavelength (blue curve, low altitude ; red curve, high altitude).

Figure. 2.3 shows the averaging kernels for the ozone density. The red line shows that the averaging kernel for ozone density equals 1 up to 42.7 km, thus below this altitude the retrieval is independent of the a priori profile. Ozone is a minor constituent in the atmosphere; due to the poor SNR of signals at higher altitudes, the sensitivity of the averaging kernel decreases. Here, the retrieval falls back to the a priori values. In a good retrieval, the difference between the forward model and the measurements, which is called the residual, should be within the uncertainty of the measurements. Figure. 2.4 shows the residual plots, which confirm that our forward model has correctly characterized the physics of the atmosphere and is capable of retrieving the quantity of interest. The OEM retrieval grid starts at 500 m and increases to 700 m at 18 km. The full width half maximum of the averaging kernel at each height is defined as the vertical resolution of the re- trieval. At lower altitudes, the averaging kernel is broad, and the retrieval resolution is close to the spacing of the retrieval grid (for this specific retrieval around 500 m). As shown in Fig. 2.5 (right panel), by increasing the altitude, the retrieval resolution decreases consequently, such that at 40 km the resolution is 2.8 km. Traditionally, the vertical resolution decreased by height as well. Figure.2.5 (left panel) shows the vertical resolution of the retrieval in both traditional methods and the OEM. At the first 2 km of retrieval the OEM provides a better retrieval resolu- 2.4. Application of the OEM to measurements from the OHP stratospheric ozone lidar 45

Figure 2.3: Averaging kernels for the ozone density for the measurements on 26 July 2017. The horizontal dashed line is a height below which the OEM retrievals is more than 80% due to the measurements. Above this horizontal cut-off as the SNR drops, the retrieval starts to fall back to the a priori profile. For clarity, the averaging kernels are only shown every 1500 m in altitude. The red line shows the summation of rows in the averaging kernel matrix at each altitude. The summation is of order unity below 42.7 km. tion, however from 14.5 km to 17 km the traditional method has a better resolution. At around 17 km both methods show the same retrieval resolution; however, the traditional resolution de- creases faster such that at 42.2 km the retrieval resolution is around 7 km. The trade-off between the retrieval resolution and the retrieval uncertainty should be considered when comparing the methods.

Having a poorer vertical resolution leads to a better (that is, smaller) retrieval uncertainty. As shown in Fig. 2.5 (right panel), the statistical uncertainty of the retrievals for the traditional method is around 12% at 15 km (where the vertical resolution is 200 m and the low altitude Rayleigh channel is used) and it decreases to less than 1% at 25 km (where the vertical resolu- tion is around 2 km and the high altitude Rayleigh channel is used). In contrast, the statistical uncertainty of retrieval in the OEM is around 10% at 15 km (where the vertical resolution is 500 m) and decreases to 2.2% at 25 km (where the vertical resolution is 700 m).

To demonstrate the mentioned trade-off in the OEM, we increased the correlation length of the a priori from 1000 m to 1500 m in the lower altitudes (below 18 km) and from 1400 m to 5500 m in higher altitudes (above 18 km). As a result, the retrieval has a poorer vertical resolution and smaller retrieval uncertainties. Assuming a higher correlation length indicates that at each altitude, the retrieved ozone density is dependent on the ozone distribution above 46 Chapter 2. Implementing the OEM to Retrieve Stratospheric Ozone Density

Figure 2.4: Residuals between the forward model and the measurements for the “on-line” and “off-line” channel (blue curves). The red line shows the uncertainty of the measurements.

and below the indicated altitude, thus, the retrieved ozone density looks smoother.

The vertical resolution and uncertainty for the traditional method as well as for the OEM with low and high correlation lengths are plotted in Fig.2.5.

In the traditional method, the relation between the final vertical resolution and the retrieval uncertainty is defined as follows: 1 3 − 2 s∝(A∆z f P0ta) (2.12)

where s is the retrieval uncertainty, A is the area of the telescope, ∆z f is the final vertical resolution, P0 is the emitted power, and ta is the acquisition time (Godin et al., 1999). Assum- ing that the traditional method has the same vertical resolution as the OEM, using the above relation we can calculate the retrieval uncertainty which corresponds to the higher vertical res- olution. Despite the difference in the vertical resolution values, at altitudes below 20 km, both the traditional method and the OEM have similar uncertainties (the difference is less than 1%). At altitudes above 20 km, assuming that the traditional method has the same vertical resolution as OEM, the retrieval uncertainty in the traditional method is calculated. Figure. 2.6 shows the comparison between OEM uncertainty and the modified traditional uncertainty for alti- 2.4. Application of the OEM to measurements from the OHP stratospheric ozone lidar 47 tudes above 20 km. As is shown in the figure, from 20 km to 35 km the difference between the uncertainties is insignificant (less than 1%) and above 35 km the difference grows to 4.5%.

Figure 2.5: Right panel: The statistical uncertainty of the OEM with correlation lengths (`s = 1000 and 1400 m is plotted (red curve) against the statistical uncertainty of the OEM with correlation lengths `s = 1400 m and 5500 m (black dotted curve). Additionally, the uncertainty of retrieval in the traditional method (blue curve) is plotted. The retrieval uncertainties in the OEM and the traditional method can be compared. Left panel: The vertical resolution of the OEM with correlation lengths (`s = 1000 and 1400 m (red curve) is plotted against the vertical resolution of the OEM with correlation lengths `s = 1400 and 5500 m (black dotted curve). The vertical resolution of the traditional method is shown as well (blue curve). The horizontal dashed line indicated the maximum height at which the retrieval is independent from the a priori.

Figure 2.7 shows our retrieved ozone density compared to the sonde measurements and the traditional retrieval. Consistent with Fig. 2.5 we have plotted the OEM retrievals for two different sets of correlation lengths. The ozonesonde measurements have better vertical resolutions compared to the DIAL measurements, albeit with larger random uncertainty. Also, the sonde profiles show more vertical structure of the ozone distribution. Compared to the traditional retrieval, the OEM can successfully catch many of these variations. As shown in Fig. 2.7 (right panel) results of a comparison between the two methods in- dicates that for higher altitudes (above 25 km) the difference between the two retrievals is insignificant. However, for lower altitudes (between 15 km to 21 km) the difference between the two methods becomes significant. Moreover, as shown in the left panel of Fig. 2.7, on this particular night, the difference between the two methods below 15 km can be as large as 60% relative to the ozonesonde, with the OEM retrieval in better agreement with the sonde 48 Chapter 2. Implementing the OEM to Retrieve Stratospheric Ozone Density

Figure 2.6: At height from 20 km to 40 km, the uncertainty of retrieval for the traditional method (assuming that it has a vertical resolution similar to the OEM vertical resolution) is plotted against the OEM retrieval uncertainty (blue curve: OEM; red curve: traditional). The horizontal dashed line indicated the maximum height at which the OEM retrieval is indepen- dent from the a priori. at most heights below 21 km. For higher altitudes the two methods agree well with the sonde measurements.

To investigate the effect of a priori profiles on retrievals, the OHP climatology and the US standard model were used to retrieve ozone density (see Fig. 2.9). The OEM retrievals resulting from these two a priori profiles as well as the traditional retrieval are plotted in the left panel of Fig. 2.9. As shown in the right panel of this figure, below 35 km the difference between the two OEM retrievals is less than 0.5%. Above this altitude, the percentage difference between the two methods reaches 2.5% which is much smaller than the retrieval uncertainty at altitudes above 35 km. Thus, the choice of a priori has a small effect on the retrievals.

The OEM provides a complete systematic and statistical uncertainty budget. Fig. 2.10 shows the uncertainty of the ozone retrieval shown in Fig. 2.7. The forward model parameters, the Rayleigh cross sections, the ozone absorption cross section, and the temperature profiles assumed for the ozone cross section contribute to the systematic uncertainty of the retrieval. Below 20 km, these uncertainties are comparable with the statistical uncertainty; however, in the higher altitudes systematic uncertainties are less than 1%. The Rayleigh-scatter cross sec- tion uncertainty, at the bottom of the retrieval, is around 7% while at higher altitudes the uncer- tainty decreases to less than 1%. These values agree with the Rayleigh-scatter uncertainty of 2.4. Application of the OEM to measurements from the OHP stratospheric ozone lidar 49

Figure 2.7: OEM ozone retrieval (red curve) from 20:07 UT to 00:15 UT on 26 July 2017 as well as the ozonesonde profile (green curve) and the traditional ozone retrieval (blue curve) are plotted. The dashed black line shows the OEM retrieval when the correlation length (łS ) became larger. The horizontal dashed line shows the cut-off below which the effect of the a priori ozone profile is small less tan 10%.

8% which is calculated in the Leblanc et al. (2016b) uncertainty budget. The ozone absorption cross section for 308 nm channel reached a maximum of 4% at the bottom of the retrieval, which is higher than the calculated uncertainty of 1% in Leblanc et al. (2016b). The uncer- tainty due to temperature is less than 0.05%. The uncertainty due to the ozone absorption cross section at 355 nm channel is negligible as well.

The ozone retrieval extends from 12 km to 70.2 km. The averaging kernel of the air density extends much higher, as the air density contributes in both back-scattering coefficients and the extinction coefficient terms in the forward model. Therefore, in air density retrievals, the maximum height of acceptable retrieval is 70.2 km. However, we show the retrievals below 42.7 km to be consistent with the ozone density retrievals. As shown in Fig. 2.11 (left panel), the relative air density profile is retrieved as well.

To validate our result, we used the nitrogen Raman spectrum at 386.7 nm. The “off-line” wavelength is transmitted to the atmosphere at 355 nm channel, and the corresponding Raman wavelength is received at 386.7 nm channel. The Raman channel is not sensitive to the aerosol contents of the atmosphere, and the wavelength is not absorbed by ozone (“off-line” Raman channel). Thus, the atmospheric back scattering and extinction terms are mostly determined by the air density. This makes the Raman “off-line” channel a good candidate for our validation. 50 Chapter 2. Implementing the OEM to Retrieve Stratospheric Ozone Density

Figure 2.8: For the night of 26 July 2017. Left panel: The percentage difference between the OEM−sonde OEM retrieval and the ozonesonde measurements in the form of: ( sonde ∗100) (blue curve); the percentage difference between the traditional retrieval and the ozonesonde measurements in traditional−sonde the form of ( sonde ∗ 100) (red curve). Right panel: The percentage difference between the OEM retrieval and the traditional retrieval (blue curve); the summation of the statistical uncertainty of the traditional and OEM retrievals (red curve).

∝ nair We can assume that N(λo f f , z) z2 . Using the above relation, the relative air density profile can be generated. The relative air density is scaled against the OEM retrieval of air density, and the percentage difference is calculated (Fig.2.11; center panel). As shown in the figure, the difference between the scaled relative air density generated from the Raman counts and the OEM relative air density is less than 10%. However, in higher altitudes (above 35 km) the difference can reach up to 50%. This difference is governed by the higher measurement noise in the Raman channel. This result provides confidence that the density retrieval is reasonable. The right panel of Fig.2.11 shows the uncertainty of the relative air density retrieval. For the air density retrieval the statistical uncertainty is small (around 0.1% at the bottom of the retrieval). The Rayleigh-scatter cross section uncertainty is small as well and the ozone absorption cross section uncertainties are negligible.

The OHP analysis employing the traditional method uses a different value of saturation correction for each wavelength. In our OEM code, we are retrieving 4 different dead times, each corresponding to one of the channels. For a priori values, we are using the provided x value which is discussed earlier in this section. As shown in Table 2.2, the retrieved dead time values for 26 July 2017 are similar to the provided x values. The only major difference is detected for the “on-line” low-altitude channel, where the x value is 4.6 ns and the retrieved 2.4. Application of the OEM to measurements from the OHP stratospheric ozone lidar 51

Figure 2.9: For the night of 26 July 2017. The left panel: the OEM retrieval using the US standard model as a priori profile (purple curve) and the OEM retrieval using the OHP clima- tology as a priori profile (red curve) are plotted. Furthermore, the traditional method retrieval (blue curve) is plotted, thus the OEM retrievals can be compared with each other and with the traditional retrieval. The right panel: Percentage difference between the OEM retrievals using the two different a priori profiles (blue curve) is plotted. This difference is with in the retrieval uncertainty. At higher altitudes (above 35 km), when the SNR drops, the difference between the two methods is less than 5%, which is smaller than the retrieval uncertainty at that height.

Dead time OEM ( ns) a priori ( ns) “on-line” high-altitude 2.78 ± 0.55 2.80 “on-line” low-altitude 5.05 ± 0.92 4.60 “off-line” high-altitude 4.60 ± 0.92 4.60 “off-line” low-altitude 2.56 ± 0.51 2.50

Table 2.2: Dead time values which were calculated for each channel on the night of 26 July 2017. value is 5.05 ns.

2.4.2 Further examples of the OEM retrieval method

Using the OEM, the retrieved profiles for the nights of 14 July and 20 July are plotted against the sonde measurements as well as the traditional ozone retrievals (Fig.2.12). The night of 14 July 2017 includes 4.5 h of measurements. The retrieval extends from 9.6 km to 40.2 km. Above 16 km, the difference between the two traditional methods and the OEM retrieval is within the statistical uncertainty of the measurements. Below 16 km the difference is about 15% with the OEM retrieval closer to the sonde measurements (Fig. 2.13). For 20 July 2017 the retrieval is the result of 4 hours of measurements. The ozone retrieval extends from 11 km 52 Chapter 2. Implementing the OEM to Retrieve Stratospheric Ozone Density

Figure 2.10: For the night of 26 July 2017. The statistical uncertainty of the OEM (blue), the Rayleigh-scatter cross section uncertainty at 308 nm (red), the ozone absorption cross section at 308 nm (orange), and the ozone absorption cross section for the 355 nm channel (purple). The horizontal dashed line shows the height below which the retrieval is independent of the a priori profile. to 36.8 km. Our results indicate that the differences between the two methods are within the retrieval uncertainty. Thus, these two additional nights help to demonstrate that the OEM can produce ozone density profiles consistent with the traditional retrievals.

2.5 Conclusion

We have introduced a first-principle OEM retrieval for stratospheric ozone profiles applicable to stratospheric DIAL lidar measurements, and tested this method using measurements from the OHP stratospheric DIAL system. The discussion of the implementation of OEM for our retrievals is summarized below.

1. The forward model used in this study is capable of providing a robust estimate of the ozone profiles for clear nights.

2. Multiple measurements channels are used. The raw (uncorrected) photocounts are used for the retrieval, and no gluing process is needed. As a result, a single ozone profile consistent with all measurements is retrieved. 2.5. Conclusion 53

Figure 2.11: Left panel: The retrieved air density (blue line) is plotted against the a priori profile (red line). Mid panel: The percentage difference between the scaled relative air density generated from the Raman channel and the OEM air density retrievals. The difference is less than 10%. Right panel: The statistical uncertainty of the OEM retrieval of air density (blue), the Rayleigh-scatter cross section uncertainty for the 308 nm channel (red), and the ozone absorption cross section in both channels (purple).

Figure 2.12: Left Panel: OEM ozone retrieval on the night of 14 July 2017 (red curve) com- pared to the ozonesonde profile (green curve) and the traditional ozone retrieval (blue curve). Right Panel: OEM ozone retrieval on the night of 20 July 2017 (red curve) compared to the ozonesonde profile (green curve) and the traditional ozone retrieval (blue curve). These cases demonstrate the high resolution of the OEM technique as evidenced by the excellent agreement around the ozone peak with the sonde measurement.

3. The OEM is applied to the OHP lidar measurements for three different nights in July 2017, all of which had coincident ozonesonde launches. Comparison with the radioson- des was good.

4. The OEM’s averaging kernels allow the contribution of the a priori relative to the mea- surements to be accessed as a function of altitude, as well as allowing better comparison with other instrument. 54 Chapter 2. Implementing the OEM to Retrieve Stratospheric Ozone Density

Figure 2.13: For the night of 14 July 2017, (a) the percentage difference between the tra- ditional method and the OEM retrieval (blue curve) plotted within the envelope of the total statistical uncertainty of the two method (red curve). The agreement between the two lidar ozone determinations are within the statistical uncertainty above 17 km. (b): The red curve is the percentage difference between the OEM retrieval and sonde measurements. The blue curve is the percentage difference between the traditional method and sonde measurements. Figures (c) and (d) are the same format as (a) and (b) for the night of 20 July 2017. 2.5. Conclusion 55

5. The OEM and the traditional method are show good agreement, and for most heights their difference is small.

6. Increasing the correlation length in the retrieval allows the vertical resolution to be de- graded and the statistical uncertainty decreased. Comparisons with the OEM retrievals at degraded resolution showed agreement to the traditional method to within the mea- surements statistical uncertainty.

7. The OEM provides a full uncertainty budget. Thus, using the OEM, for each individual retrieved profile both statistical and systematic uncertainties are calculated. The system- atic uncertainties are compared with the uncertainty budget for the traditional method given by (Leblanc et al., 2016a) and are similar.

Currently we are working on a retrieval which can use measurements from both the OHP tro- pospheric and stratospheric lidars which will allow us to retrieve ozone profile from just above the boundary layer throughout the stratosphere. Also, we plan to include the Raman measure- ments into our forward model, allowing the retrieval of the ozone profiles in the presence of strong aerosol layers and thin clouds. Also, we are planning to apply our OEM retrieval to the last three decades of OHP measurements. Applying the OEM to the entire OHP lidar ozone profile database will provide an improved statistical evaluation of the differences between the traditional and the OEM methods, as well as allowing improved ozone estimates in the upper troposphere and lower stratospheric region. 56 Chapter 2. Implementing the OEM to Retrieve Stratospheric Ozone Density

Bibliography

Eriksson, P., Jimenez, C., and Buehler, S.: Qpack, a general tool for instrument simulation and retrieval work, J. Quant. Spectrosc. Radiat. Transfer, 91, 47–64, 2005.

Godin, S., Carswell, A. I., Donovan, D. P., Claude, H., Steinbrecht, W., McDermid, I. S., McGee, T. J., Gross, M. R., Nakane, H., Daan, Swart, P. J., Bergwerff, B. B., Uchino, O., von der Gathen, P., and Neuber, R.: Ozone differential absorption lidar algorithm intercom- parison, Appl. Opt., 38, 6225–6236, 1999.

Godin-Beekmann, S., Porteneuve, J., and Garnier, A.: Systematic DIAL lidar monitoring of the stratospheric ozone vertical distribution at Observatoire de Haute-Provence (43.92[de- gree]N, 5.71[degree]E), J. Environ. Monit., 5, 57–67, 2003.

Krueger, A. J. and Minzner, R. A.: A midlatitude ozone model for the 1976 US Standard Atmosphere, Journal of Geophysical Research: Atmospheres (19842012), 81, 44774481, 1976.

Leblanc, T., Sica, R. J., van Gijsel, J. A. E., Godin-Beekmann, S., Haefele, A., Trickl, T., Payen, G., and Gabarrot, G.: Proposed standardized definitions for vertical resolution and uncertainty in the NDACC lidar ozone and temperature algorithms Part 1: Vertical resolu- tion, amt, 9, 4029–4049, 2016a.

Leblanc, T., Sica, R. J., van Gijsel, J. A. E., Godin-Beekmann, S., Haefele, A., Trickl, T., Payen, G., and Liberti, G.: Proposed standardized definitions for vertical resolution and uncertainty in the NDACC lidar ozone and temperature algorithms Part 2: Ozone DIAL uncertainty budget, amt, 9, 4051–4078, 2016b.

Malicet, J., Daumont, D., Charbonnier, J., Parisse, C., Chakir, A., and Brion, J.: Ozone UV spectroscopy. II. Absorption cross-sections and temperature dependence, J. ATMOS. CHEM., 21, 263–273, 1995.

McDermid, I. S., Godin, S. M., and Walsh, D.: Lidar measurements of stratospheric ozone and intercomparisons and validation, Appl. Opt., 29, 4914–4923, 1990. BIBLIOGRAPHY 57

Nicolet, M.: On the molecular scattering in the terrestrial atmosphere: An empirical formula for its calculation in the homosphere, Planet. Space Sci., 32, 1467–1468, 1984.

Orphal, J., Staehelin, J., Tamminen, J., Braathen, G., De Backer, M., Bais, A., Balis, D., Barbe, A., Bhartia, P. K., Birk, M., et al.: Absorption cross-sections of ozone in the ultraviolet and visible spectral regions: Status report 2015, J. Mol. Spectrosc., 327, 105–121, 2016.

Pelon, J. and Megie,´ G.: Ozone monitoring in the troposphere and lower stratosphere: Evalua- tion and operation of a ground-based lidar station, J. Geophys. Res. Oceans, 87, 4947–4955, 1982.

Rodgers, C. D.: Inverse methods for atmospheric sounding: theory and practice, vol. 2, World scientific, 2000. Chapter 3

Improved ozone UTLS DIAL measurements

3.1 Introduction

The upper troposphere and lower stratosphere (UTLS) region extends from about 6 km to 25 km in height and plays a significant role in the atmospheric climate system. In this region of the atmosphere, even small changes in temperature and in the distribution and concentration of greenhouse gases can result in large changes in atmospheric radiative forcing, which can trigger climate change IPCP (2007); Logan (1985). Ozone in the upper troposphere acts as the third largest greenhouse gas contributing to the radiative forcing of climate change Ramaswamy et al. (2001); IPCP (2007). The ozone distribution in the UTLS is the result of transport mechanisms and photochemical reactions. Because of stratospheric tropospheric exchange, large spatial and temporal variability can be observed in the UTLS Forster et al. (1997). In many studies on the UTLS ozone, satellite-borne instruments are used. In limb-viewing instruments, the elevation angle of the line-of-sight varies during the measurements. As a result, limb sounders can provide good vertical resolution (about 2 km to 4 km). However, at lower altitudes (lower troposphere), the atmosphere becomes nearly opaque, and the limb- viewing instruments have difficulties measuring trace gases. Nadir-viewing instruments can provide measurements in the lower troposphere, but their vertical resolution is limited (about

58 3.1. Introduction 59

6 km to 7 km). Occultation instruments use the Sun or other stars as the source of radiation, and they can obtain measurements with higher vertical resolution (about 1 km to 2 km). Solar occultation instruments are restricted by the number of sunsets and sunrises they encounter in one orbit, while stellar occultation instruments are limited by the weakness of the stellar source compared to the Sun. The combination of measurements from different geometrical- based satellite instruments has been used to measure ozone density.

The OEM methodology we describe will allow improved estimates of UTLS ozone and the associated random and systematic uncertainties, as well as provide averaging kernels for the lidar measurements. The availability of averaging kernels will improve future intercomparisons between ground-based, balloon-borne and space-based ozone measurements.

Differential Absorption Lidar (DIAL) systems provide ozone measurements with high ver- tical and temporal resolutions. For example, observatories such as the Canadian Network for the Detection of Atmospheric Change (CANDAC) Polar Environment Atmospheric Research Laboratory (PEARL) in Eureka, Ma¨ıdo observatory in Reunion Island, the Observatoire de Haute Provence (OHP) in France, and the NASA Table Mountain Observatory (TMO) in the United States are equipped with both tropospheric and stratospheric lidars. At the Eureka ob- servatory, the tropospheric lidar system makes measurements from 0.5 km to about 8 km in altitude and the stratospheric lidar system operates from about 4 km to 35 km (??). At the Ma¨ıdo observatory, the tropopspheric DIAL makes measurements from 6 km to 16 km, and the stratospheric DIAL operates in the 13 km to 38 km region Baray et al. (2013). At the OHP observatory, the tropospheric DIAL system operates from 2.5 km to about 14.5 km, and the stratospheric DIAL operates from about 10 km to 45 km Gaudel et al. (2015); Godin-Beekmann et al. (2003). At the TMO, the tropospheric DIAL system obtains measurements from 3 km to 18 km, and the stratospheric DIAL system from 10 km to 40 km Megie et al. (1985); Merienne´ et al. (2001). Although these systems can produce satisfactory ozone profiles in their overlap- ping region (from tropospheric lidar to stratospheric lidar), the uncertainty of merging is not well defined. Providing a single ozone profile with a full uncertainty budget using both sets of measurements can significantly improve our measurements of ozone in the UTLS Holton et al. (1995); Stohl et al. (2003); Cohen et al. (2018).

Here we apply the Optimal Estimation Method (OEM) to tropospheric and stratospheric 60 Chapter 3. Improved ozone UTLS DIAL measurements

DIAL measurements. Measurements from these two systems are simultaneously used by the retrieval to obtain a single ozone profile. Using the OEM there is no need to “merge” or “glue” Level 0 profiles. Moreover, the input measurements can be in different units with different mea- surement grids (for example a mix of analog and digital measurements). Additionally, a full uncertainty budget, including both the systematic and statistical uncertainties, is calculated for each individual profile. The OEM also provides averaging kernels of the retrievals, which al- lows comparison of the profiles with other measurements which can account for differences in vertical resolution, such as when compared to space-based measurements. Other atmospheric and systematic parameters such as air density, the dead time of the system, and the background counts can be retrieved along with ozone profiles. The application of OEMs to aerosol li- dar measurements, Rayleigh scatter temperatures, and Raman scatter water vapour retrievals has been studied and discussed in detail Povey et al. (2014); Sica and Haefele (2015, 2016). In addition, we have recently demonstrated an OEM for DIAL stratospheric ozone retrievals (Farhani et al., 2018), which we will now expand to include measurements from tropospheric ozone DIAL systems.

In this chapter, focusing on the UTLS region, we show a first principle OEM to retrieve a single ozone profile by using both tropospheric and stratospheric DIAL measurements directly from the raw (Level 0) measurements using the lidar equation as the forward model. In Sec- tion 3.2, pre-processing steps prior to applying the traditional DIAL algorithm, as well as the OEM, are discussed. Moreover, the state vectors and the b parameter quantities are defined and a brief overview of the lidar’s specifications is given. In Section 3.3 results of the OEM retrieval, using both tropospheric and stratospheric lidar measurements, are discussed in detail. In this Section, we also show our results of comparison between the ozonesonde profiles and our retrievals. Details of how to apply our method to a standalone tropospheric DIAL are given in the 3.4. Section 3.5 is the summary of the chapter, and in Section 3.6 we discuss our future plans. 3.2. METHODOLOGY 61

3.2 METHODOLOGY

In the DIAL system, two wavelengths are simultaneously transmitted to the atmosphere. One of the emitted wavelengths is strongly absorbed by the constituent of interest (called the “on-line” wavelength) and the other is weakly absorbed (called the “off-line” wavelength). For ozone measurements, selecting a wavelength pair depends on the altitude range of the measurements. For most studies, the ultraviolet (UV) spectrum is the most efficient spectral region. A pair of wavelengths with strong UV absorption is needed to detect the small amount of ozone which resides in the troposphere. However, for stratospheric ozone measurements, choosing a laser with a longer wavelength that can reach higher altitudes in the stratosphere is the main concern Megie et al. (1985); Browell (1989); Papayannis et al. (1990).

The traditional analysis method for ozone number density uses the derivative of the ratio between the “on-line” and “off-line” channels to calculate the ozone number density no3(z).A detailed discussion on the tropospheric and stratospheric ozone retrievals can be found in An- cellet et al. (1989); Godin et al. (1999); McDermid et al. (1990); Godin-Beekmann et al. (2003); Leblanc et al. (2016). In the traditional analysis, some corrections are applied to the raw lidar measurements, for example background counts should be removed. In many systems this re- quires including the effects of signal-induced-noise (SIN). Any corrections due to nonlinearity of the counting system (because of saturation) should also be applied to the raw counts. Finally, the signals from different channels need to be merged to form a single measurement profile. This corrected count profile is then used to calculate the ozone number density profiles. With the OEM, a forward model encapsulates the geophysical properties and instrumental charac- teristics of the system, and our OEM retrieval uses the raw (Level 0) measurements from all available channels. Unlike the traditional method, the OEM does not require corrections to the raw measurements. Furthermore, the measurements from different channels are not glued (or merged), but are input directly into the OEM routine as the measurement input vector. The dead time of the system and the background values for the “on-line channel and the “off-line channel are part of the state vector, while the overlap function is a model parameter whose con- tribution to the retrieved ozone profile is assessed in the uncertainty budget. The dead times, backgrounds, and lidar constants are simultaneously retrieved along with the ozone number 62 Chapter 3. Improved ozone UTLS DIAL measurements density and air number density profiles. A comprehensive explanation of the OEM can be found in Rodgers (2000); a brief description of the OEM follows below.

The OEM is an inverse modeling technique which is based on Bayes’ theorem. In the OEM a forward model is defined as the relation between the measurement vector y = (y1, y2, ..., yn), and the state vector x = (x1, x2, ..., xn). The forward model is:

y = F(x, b) +  (3.1) where b are the forward model parameters, which are assumed to be known, and  is the measurement noise.

We use the lidar equation as the forward model, where the raw counts are the measurements.

The lidar equation for unsaturated counts, Ntrue, is:

C(λ )O(z) Z ∞ X N (z, λ ) = i β(λ , z) exp[−2 [σ (λ, T(z))n (z)+α(λ, z)+ σ (λ)n (z)]dz]+N (z, λ ) true i z2 i O3 O3 e e b i 0 e (3.2)

where Ntrue(z, λi) is the number of backscattered photons. C(λi) is the lidar constant, which contains the area of the receiving telescope, the total efficiency of the lidar system, and en-

ergy of the scattered photon. The geometrical overlap is O(z), and β(λi, z) are the atmospheric backscattering coefficients which includes both molecular and aerosol terms. The first term

inside the integral corresponds to ozone absorption in which σO3 (T(z), λi) is the ozone absorp-

tion cross section, which depends on atmospheric temperature, and nO3 (z) is the ozone number density. The second term of the integral, α(λ, z) contains the extinction coefficient which is P the sum of the extinction due to molecules and particles, and the last term e σe(λ)ne(z) is the

extinction by other absorbers. For ozone studies, the most common interfering gases are SO2 ,

NO2 and O2. The effect of O2 is only considered when the selected “on-line” laser wavelength is shorter than 294 nm Fally et al. (2000); Merienne´ et al. (2001). In the case of heavy volcanic eruption, SO2 and NO2 can significantly affect ozone retrievals Heath et al. (1983). However, in most cases, for both stratospheric and tropospheric ozone studies the effect of these gases in final ozone retrievals is negligible. Thus, the last term of integration is typically neglected Godin-Beekmann et al. (2003). 3.2. METHODOLOGY 63

The background counts are written as Nb(z). In the presence of SIN, the background is fitted to an exponential function of the form:

Nb(z) = a exp(−bz) + c (3.3)

where a, b, and c are coefficients of the fit, which in the traditional method are determined ana- lytically, but are retrieved in our OEM retrieval using the analytic values as a priori coefficients Hunt and Poultney (1975). When the intensity of the backscattered signal is high, the counting system can be affected by saturation. This saturation can result in an observed count rate which is less than the true

count rate (Ntrue). For a paralyzable detector, true counts are related to the observed counts

Nobs as follows:

Nobs = Ntrue exp(−κNtrue) (3.4) and, for non-paralyzable detectors, the following equation can be used:

Ntrue Nobs = (3.5) 1 + κNtrue where κ is the dead time of the detecting system. For the OEM retrieval the value of the dead time for each channel is retrieved.

3.2.1 Implementing the OEM for the OHP lidars

Knowledge of the measurement vector y and its covariance matrix, Sy, along with a priori values of the state vector, xa, and its associated covariance matrix, Sa, enables the OEM to calculate an optimal a posteriori state by minimizing a cost function with respect to x given by: T −1 T −1 Cost = (y − Kx) Sy (y − Kx) + (x − xa) Sa (x − xa) (3.6)

dF where K = dx , the Jacobian matrix, is the linearisation term for the nonlinear forward model. In the OEM it is assumed that the measurement noise is described by a normal distribution, but lidar photocount measurements follow a Poisson distribution. Since at most heights the number 64 Chapter 3. Improved ozone UTLS DIAL measurements of photocounts is large, the two distributions are indistinguishable. Thus, we assumed that the photon counts are distributed normally. This assumption causes the residuals to become biased at very low count rates, but this bias occurs at altitudes far above where the retrieval is valid as defined by the response function, the area of the averaging kernel matrix which is of order unity, when the retrieval depends fully on the measurement rather than the a priori profile.

As our forward model is nonlinear, an iterative numerical method is used. For our prob- lem the Levenberg-Marquardt iteration is a suitable numerical method. Then, the optimized solution for the state vector x is given as:

−1 T −1 T −1 −1  xi+1 = xi + [(1 + γi)Sa + Ki SyKi] [Ki Sy (y − F(xi, b)] − Sa (xi − xa) (3.7)

where i is the iteration term, and γi is a damping factor for the iteration, which is chosen at each step to minimize the cost function. As suggested by Fletcher Fletcher (2013) if the value of the cost function increases in a step, γi will increase by a factor of 10, and if the value of the cost function decreases in a step, γi will decrease by a factor of 2. The iteration stops when the cost function decreases to a value much smaller than the number of measurements. There are other criteria which result in ceasing the iteration. Further details can be found in Rodgers (2000).

To understand how measurements and a priori profiles contribute in the final retrievals, an averaging kernel can be used. The relation between the retrieved state and the true state is described by the averaging kernel of the retrieval. The averaging kernel is calculated as:

dx A = b = [KT S−1K + S−1]−1KT S−1K (3.8) dx y a y

The retrieved quantity (bx) can be written as follows:

bx = (I − A)xa + Ax + r (3.9)

where r is the retrieval uncertainty and I is a unity matrix. A perfect retrieval, in the sense all the information comes from the measurement with no effect from the a priori state, has averaging kernels equal to one, where the first term of the above equation becomes zero. The 3.2. METHODOLOGY 65

width of the averaging kernel gives the resolution of the retrieval at each height, here defined as the Full Width Half Maximum (FWHM) of each averaging kernel. In order to find the state vector (from Eq. 3.7) the following quantities should be known: the measurements and their covariances, the a priori profiles, the a priori profile’s covariance, and the model (b) parameters. The b parameters are quantities in the forward model that are not being retrieved, because they are either well-known or retrieving them is not possible. The uncertainty associated with the retrieval due to the b parameters is calculated after the last iteration of the solution. We used the Qpack package to perform the retrievals. Qpack is a free software package which written in MATLAB, and is part of the ATMOS package developed for retrieval of geophysical quantities from passive instruments in the millimeter and sub-millimeter wavelength regions. Here, we have developed our own forward model, and we have also calculated the Jacobians ourselves. After assigning the a priori profiles, b parameters and their associated uncertainties, we then use the Qpack OEM solver (oem.m) to retrieve the state vector using the Marquardt-Levenberg technique. Details of the Qpack software are given by Eriksson et al. Eriksson et al. (2005). Here we retrieve the ozone density profile, relative air density, dead time values, and back- ground counts. Overlap functions, ozone cross sections, and Rayleigh scattering cross sections are considered as b parameters in the forward model. Below, we discuss our choices of a pri- ori profiles and b parameter values. The covariance matrices associated with the measurements and a priori profiles are discussed as well, and these values are summarized in Table 3.1. In photon counting mode, when the signal is linear, the measurements statistical uncer- tainty follows a Poisson distribution, and the number of counts at each altitude represents the measurement’s variance at that height. There is no correlation between the digital counts in dif- ferent layers of the atmosphere, so the off-diagonal elements of the measurement’s covariance matrix are zero. However, for the OHP lidars, analog measurements do not follow Poisson distributions. Calculating the measurement variance for each measurement point requires se- lecting n points before and after the specified point, and then fitting a straight line to these 2n+1 points, which is then removed. Next, the residual variance is calculated. For our measurements, we tried different values for n, and n = 3 provided the best fit for our measurements. In order to determine the background counts in both the tropospheric and stratospheric 66 Chapter 3. Improved ozone UTLS DIAL measurements measurements, the mean of the counts above a specific height is calculated and used as the a priori for the “off-line” channels, since SIN is negligible in these channels. The variance of the background counts divided by the number of bins in the selected region is the uncertainty for the background a priori value. For “on-line” channels where SIN is present, an exponential function is fitted to the signal and the coefficients of the fit are used as a priori values. An uncertainty of 20% is assigned to these coefficient a priori values. The altitude above which the background counts are determined is different for tropospheric and stratospheric lidars. Also, as the laser power in the “online” channel is about 2 times stronger than the laser power in the “off-line” channel, the effect of SIN in the “on-line” channel is more pronounced. The values we chose for the OHP lidars measurements are shown in Table 3.1.

We retrieve the logarithm of ozone density, q(z) = ln{no3 (z)}, so in Eq. 3.2, no3 is replaced by eq, as is commonly done for atmospheric retrievals (e.g. Deeter et al. (2007) and Sica and Haefele (2016)). Retrieving the logarithm of the number density is numerically more favourable to show small changes in large numbers. The U.S standard model is used for the a priori ozone profile (Krueger and Minzner, 1976). An uncertainty of 20% is assigned to this profile.

We retrieve the air density for both tropospheric and stratospheric measurements. However, below 15 km the air density profile retrieved is in fact a convolution of air density and aerosol load. Mass Spectrometer Incoherent Scatter Radar (MSIS) total density profiles are used as a priori profiles, and an uncertainty of 20% is assigned to it (Hedin, 1991). To generate a full length a priori covariance matrix for both air and ozone density profiles at altitudes below 12 km, a tent function with a correlation length of 300 m was used. At altitudes above 12 km the correlation length was increased to 900 m. This choice of correlation length is because above 12 km, the stratospheric lidar measurements have the most weight and the retrieval grid for these measurements starts at 300 m. Using the forward model, the a priori of the lidar constants for both tropospheric and stratospheric lidars are estimated. We assign a 10% uncertainty to the a priori of the lidar constants to account for changes with time of laser power, atmospheric transmission, and system efficiency.

The overlap function, Rayleigh cross sections, and ozone absorption cross sections are assumed as b parameters. Although these parameters are not being retrieved, the uncertainty 3.2. METHODOLOGY 67

Parameter Value Standard Deviation Stratospheric lidar measurements measured Poisson statistics Tropospheric lidar measurements (digital measured Poisson statistics channels) Tropospheric lidar measurements (analog measured 3-point running standard de- channels) viation Retrieved quantities a priori profiles Standard Deviation Ozone density U.S standard model 20% Air density MSIS 20% Deadtime empirical fitting 20% Background for stratospheric measure- mean above 80 km σ above 80 km ments (“off-line”) Coefficients of SIN for stratospheric mea- empirical fitting above 80 km 20% surements (“on-line”) Background for tropospheric measure- mean above 20 km σ above 20 km ments (“off-line”) digital channel Background for tropospheric measure- mean above 12 km σ above 12 km ments (“off-line and “on-line”) analog channel Coefficients of SIN for tropospheric mea- empirical fitting above 12 km 20% surements (“on-line”) digital channel Lidar constants estimate from FM 20% Forward model parameters Sources Standard Deviation Rayleigh-scatter cross section Eberhard Eberhard (2010) 2% Ozone absorption cross section BDM Malicet et al. (1995) 2% Overlap function available at Dataset Cita- 10% tion 3.6.1

Table 3.1: Values and associated uncertainties for the retrieved and forward model parameters. associated to them contributes to the final uncertainty budget. The assigned values for the overlap function, Rayleigh cross sections and ozone cross section along with their standard deviations are listed in Table (3.1).

3.2.2 Description of the Lidars

The OHP Observatory (44◦N, 6◦E) has made routine measurements of ozone in the free tropo- sphere and stratosphere for the last three decades using two lidar systems, a tropospheric DIAL and a stratospheric DIAL. The transmitter for the tropospheric system uses the fourth harmonic of a Continuum Nd:YAG laser (266 nm) frequency shifted by Raman Stimulated Scattering in 68 Chapter 3. Improved ozone UTLS DIAL measurements a D2 high pressure cell Ancellet and Beekmann (1997). The DIAL measurement makes use of the 1st and 2nd Raman Stokes lines at 289 nm (the “on-line” wavelength) and 316 nm (the “off-line” wavelength). Backscattered photons are collected by a Cassegrain telescope which is equipped with a 80 cm mirror. For the spectral separation of the two backscatter signals the collected signal is passed from the telescope to a spectrometer.The signals are detected by a photomultiplier tube (PMT). The system configuration is bi-axial, and the distance between the laser and the telescope axes is 0.5 m. The overlap, O(z), is significant from the surface to about 4 km Halldorsson and Langerholc (1978). The stratospheric lidar system uses an XeCl excimer laser at 308 nm, with a repetition rate of 100 Hz. This laser has an output energy of about 200 mJ for the “on-line” channel, while the “off-line” wavelength at 355 nm is generated by the third harmonic of a continuum Nd:YAG laser with an energy of 50 mJ at 50 Hz. The backscattered signal is collected by four Newtonian telescopes, each with diameter of 0.5 m. The collected signal is sent to a spectrometer which separates the signal into four wavelengths. Two of these correspond to the emitted Rayleigh signals at 308 nm and 353 nm. The other two correspond to the Nitrogen Raman shifted spec- trum at 331.8 nm and 386.7 nm, respectively. The Rayleigh signals are separated further into high- and low-gain photomultiplier channels. Hence, in total 6 photocount profile measure- ments are obtained. Further details on the OHP tropospheric and stratospheric lidars can be found in Ancellet and Beekmann (1997); Gaudel et al. (2015); Godin-Beekmann et al. (2003).

3.3 OEM ozone retrieval in the free troposphere and strato- sphere

In this section we present the result of combining the two lidar measurements to retrieve a single ozone profile. We choose measurements from 12 July 2017 as this night has both clear skies and coincident measurements from the NDACC-LAVANDE intercomparison campaign. The results for the nights of 14 and 26 July 2017 are presented as well (see Table. 3.2). Our first example retrieval will be from the night of 12 July 2012, where the tropospheric lidar operated from 2049 to 2357, and the stratospheric lidar operated from 2106 to 0142 (all local time). An 3.3. OEM ozone retrieval in the free troposphere and stratosphere 69

Dates 12 July 2017 14 July 2017 26 July 2017 Tropospheric lidar 2049 – 2357 2350 – 0219 2140 – 2237 Stratospheric lidar 2106 – 0142 2151 – 0221 2007 – 0016 Ozone sonde 2153 2348 2133

Table 3.2: Measurement periods for the tropospheric and stratospheric lidars systems and launch times for the ozonesondes. ozonesonde was launched at 2153 from the OHP station and the tropopause height was at 14.7 km.

For the tropospheric lidar system, the native resolution of measurements is 7.5 m for the analog channels and 30 m for the digital channels. For the stratospheric lidar system, the native resolution of measurements for all six channels is 150 m. Our retrieval starts at 2.6 km with a resolution grid of 150 m. At 11 km the retrieval grid changes to 500 m, and at 21 km it changes to 1500 m, and finally, at 25 km height, it becomes 1700 m. We chose these retrieval grids to be closer in vertical resolution to traditional retrieval grids.

The averaging kernels calculated for the retrieval are shown in Fig. 3.1. The shape of the averaging kernels define the sensitivity of the retrieval to the true state. As shown in Eq. 3.9, when the averaging kernel equals 1, the retrieval is sensitive only to the measurements. The vector area of the averaging kernels is defined as Au, where u is a unity vector. When the area is close to 1, the retrieval is mostly independent of its a priori value. In our retrieval the averaging kernel has an area of about 1 up to 42.2 km, indicating the retrieval is mostly independent of the a priori profile. At about 11 km the stratospheric measurements are added to the retrieval. The stratospheric measurements have much lower noise at these heights than the tropospheric measurements. This rapid change in the measurement variance results in the spike seen in the response function. This behaviour of the response functions is well known in satellite OEM retrievals, and, if severe, can be mitigated by adjusting the covariances in the transition region. For our retrievals the perturbation region is narrow in height and does not cause any significant variation in the ozone profile retrieved in the UTLS, so we did not adjust the retrieval. We plan to investigate ways to smooth the response function in the hand-off region as we improve our retrieval.

The residual plots, which show the difference between the forward model and the actual 70 Chapter 3. Improved ozone UTLS DIAL measurements

Figure 3.1: Averaging kernels for tropospheric-stratospheric ozone measurements. The aver- aging kernels are only shown every 450 m for lower altitude (from 2.5 km to 11 km height), and every 1500 m in higher altitudes for clarity. As is shown by the response function (red curve), the area of the averaging kernel matrix has a small spike at 11 km, when the stratospheric ozone measurements are included.

measurements, for both the tropospheric and stratospheric lidar are shown in Fig. 3.2. The four plots on the left are the residuals for the stratospheric measurements, and the four plots on the right are residuals for the tropospheric measurements. As shown in the figure, for both low- altitude and high-altitude channels the forward model can successfully encapsulate the physics of the atmosphere and the characteristics of the lidars, and, up to 50 km, the difference between the forward model and the actual measurements is less than 5%.

Figure 3.2: The percentage difference between the forward model and the actual measurements are shown in blue. The statistical uncertainty is plotted in red. The four plots at the left are stratospheric forward model residuals, and the four plots at the right are the tropospheric forward model residuals. For digital counting systems, the Poisson distribution is appropriate and the variance of the measurements at each altitude is the number of photons at that altitude. However, the output signals of the analog channels do not follow a Poisson distribution and to find the variance a 3-point running filter is used. As a result the red line (which indicates the noise of measurements) for the analog channels is more structured than for the digital channels. 3.3. OEM ozone retrieval in the free troposphere and stratosphere 71

The ozone profile for the night of 12 July 2017 is retrieved from 2.6 km to 42.2 km altitude. In Fig. 3.3 the OEM retrieval is plotted against the traditional stratospheric and tropospheric retrievals. The traditional ozone retrieval starts from 2.5 km and extends to 14.5 km, and the traditional stratospheric ozone retrieval starts at 11 km and extends to about 42.2 km. The tropopause height on this night is at 14.7 km. For comparison purposes the ozonesonde profile which starts from the ground and goes up to 33 km is shown as well. On this night of measure- ments, the ozonesonde balloon was released at 2153 from the station (44◦N, 5.8◦E). During the time of fly it drifted southeastward, such that 1.5 hours later, at the altitude of 33 km, its location was (43.6◦N, 6◦E).

To demonstrate how the OEM retrieval performs in the region where the tropospheric mea- surements are merged with the stratospheric measurements, we consider the retrievals in the region between 5 km to 18 km (Fig. 3.3). At around 6 km altitude the OEM retrieval shows a decrease in ozone not apparent in the traditional method or by the ozonesonde. However, the vertical resolution of the OEM retrieval at this altitude is half that of the traditional retrieval, with a subsequent doubling of the statistical uncertainty (Fig. 3.6). Thus, over the 5.5 km to 6.5 km altitude region the lidar ozone profiles are similar to within the measurement uncer- tainty, and less than the ozonesonde consistently up to about 15 km altitude. In the hand-off region between the tropospheric and stratospheric systems, the OEM retrieval is in general closer to the ozonesonde than the traditional method applied individually to both lidars, except around 11 km. Here there is a bump in the traditional method for the tropospheric system not seen in the OEM retrieval. At this altitude the traditional method has a vertical resolution of about 800 m and like the OEM a statistical uncertainty of about 5%. As was true for the feature around 6 km, the different in lidar ozone values is within the statistical uncertainty when the ozone values in the region of the bump are averaged over 800 m.

The vertical resolution of the OEM retrieval is calculated from the full width at half max- imum (FWHM) of the averaging kernel at each altitude. The vertical resolutions as well as the statistical uncertainties of the OEM and the traditional retrievals are plotted in Fig. 3.4. In the free troposphere, at an altitude of 2.5 km, the vertical resolution for the OEM retrieval is 150 m, increasing to 300 m at 11 km. The traditional vertical resolution starts at 150 m as well, but designed to grow faster such that at 11 km the vertical resolution is 1000 m (Fig. 3.4, 72 Chapter 3. Improved ozone UTLS DIAL measurements

Figure 3.3: The OEM retrieval (red curve) compared to the traditional calculation of ozone using the OHP lidar systems. The tropospheric lidar starts at 2.5 km and extends upward to 14.5 km (blue curve). The ozone profile measured by the stratospheric lidar system (black curve) overlaps with the profile retrieved from the tropospheric lidar system in the UTLS. In this region, the OEM retrievals smoothly transition from relying primarily on the tropospheric lidar measurements to the stratospheric measurements.

right panel). The trade-off is that the uncertainty of the retrieval in the traditional method is smaller, so that at 11 km it is 4.5% as opposed to the OEM retrieval which has a larger uncer- tainty of 7.5%. At 11 km where the stratospheric measurements are added, the OEM vertical resolution is 300 m and gradually increases up to 600 m at 14.5 km, whereas, in the traditional tropospheric method, the retrieval resolution increases to 1900 m at the same height. At 21 km altitude, the vertical resolution is 1500 m, while at 25 km it is 1700 m. The vertical resolution does not change until 40 km, where due to the rapid drop in SNR it increases to 2000 m. The percentage difference between the OEM retrieval and the ozonesonde measurements is shown in Fig. 3.5. For most heights the difference between the OEM retrieval and the sonde measurements in within the uncertainty of the two profiles. At 10 km, the difference between the sonde and the OEM retrieval is almost 30%. Above this altitude, and in the UTLS, the difference between the two profiles is less than 10% The systematic and statistical uncertainties for the retrieved ozone profile for this night are shown in Fig. 3.6. The systematic uncertainties due to Rayleigh cross sections, the ozone absorption cross sections, and the overlap function are the b parameters in the forward model which contribute to the uncertainty of the ozone retrieval. 3.3. OEM ozone retrieval in the free troposphere and stratosphere 73

Figure 3.4: Left panel: The statistical uncertainty of the OEM retrieval (red curve) is plotted against the statistical uncertainty of the traditional retrievals. The uncertainty of the retrieval for the stratospheric and tropospheric lidar systems respectively are shown in black and blue. Right panel: The vertical resolution of the OEM retrieval is shown in red. The vertical resolution of the traditional calculation from the tropospheric lidar system is shown in blue, while the vertical resolution of the retrieved profile produced from the stratospheric lidar system is shown in black.

Figure 3.5: The percentage difference between the OEM retrieval and the ozonesonde mea- surements (blue curve) is plotted within the total statistical uncertainty of the OEM retrievals plus the ozonesonde measurement (red curves).

The contribution of the Rayleigh-scatter cross section uncertainty to the ozone retrievals for both the tropospheric and stratospheric lidars is less than 1%, consistent with the calculations given by Leblanc et al. Leblanc et al. (2018). The ozone absorption cross section for the 289 nm channel is about 5% which is close to the 7% uncertainty calculated by Leblanc et al. (2016). For stratospheric measurements, the ozone uncertainty has its maximum of 4% at the bottom of retrievals, which is higher than the calculated uncertainty of 2% in uncertainty budget of Leblanc et al. (2016). This difference is due to the OHP tropospheric DIAL having a larger wavelength separation than that in the NDACC LWG calculation. The uncertainties due to the overlap function is 5% at the bottom of the retrieval and, at 74 Chapter 3. Improved ozone UTLS DIAL measurements

Figure 3.6: Uncertainty budget on the night of 12 July 2017. The statistical uncertainty of the retrieval (blue), the Rayleigh-scatter cross section uncertainty at 308 nm (dashed line red), the Rayleigh-scatter cross section uncertainty at 289 nm (dashed line yellow), the ozone absorption cross section at 308 nm (dashed line purple), the ozone absorption cross section for the 289 nm channel (dashed line green), and the overlap function for the 289 nm channel (dashed line light blue) all contribute to the budget. The horizontal dashed line shows the height below which the retrieval is independent of the a priori profile.

10 km it drops to about 1%. The uncertainty due to the ozone cross sections at 316 nm and 353 nm are negligible and are not shown in this plot. Also, the uncertainty on the retrieved ozone profile due to the temperature uncertainty of the ozone cross section is negligible as well. Ozone density profiles are also retrieved for the nights of 14 July 2017, and 26 July 2017, each of which also had coincident ozonesonde measurements. On July 14 2017 the tropopause height is 12.5 km. The traditional retrievals of tropospheric and stratospheric lidars, at alti- tudes between 10 km to 15 km, are not consistent with each other (Fig. 3.7). The percentage difference between the traditional tropospheric and stratospheric ozone profiles in this region reaches its maximum of 33% at a height of 11.8 km. The OEM retrieval, similar to the night of 12th July, smoothly hands off from one lidar’s measurements to the other one, and in this case is closer to the traditional stratospheric DIAL measurement. Also, the statistical uncertainty of the OEM retrieval reaches its maximum at about 12 km. The figure also shows the OEM re- trieval compared with the sonde measurements. The ozonesonde was released from the station (44◦N, 5.8◦E) and it flew to the southeast. At its maximum height, the ozonesonde was located 3.3. OEM ozone retrieval in the free troposphere and stratosphere 75 at (43.5◦N, 6.3◦E), and was not more than 50 km away from the OHP station. Similar to 12 July 2017, the percentage difference between the OEM and the sonde measurements is within the two profiles uncertainty.

On the night of 26 July 2017, the two ozone profiles calculated by the traditional method are inconsistent in the region from 12 km to 14 km are inconsistent with each other in the region of the tropopause (13.3 km). The OEM retrieval can smoothly transition from the tropospheric measurements to the stratospheric measurements (Fig. 3.8). Although, the OEM and the tra- ditional analysis in the lower troposphere are match well (their difference is about 2%), the sonde profile is far from the two retrievals, and the difference between the sonde measurements and the lidar measurements (both the OEM and the traditional analysis) is greater than 60%. The ozonesonde, similar to the other nights, was released from the station, but comparing to the other nights it moved slightly farther toward the south, such that at its maximum height its location was (43.1◦N, 5.8◦E). Thus, the sonde was within approximately 100 km of its launch point.

Figure 3.7: OEM ozone-profile retrieval on 14 July 2017 (red curve). The tropospheric tradi- tional retrieval (blue curve) extends from 2.5 km to 15 km, while the stratospheric traditional retrieval (black curve) extends from 10 km to 43 km. At the region where the tropospheric and stratospheric lidar measurements overlap, the OEM can smoothly makes a transition from one lidar system’s measurements to the other system’s measurements. 76 Chapter 3. Improved ozone UTLS DIAL measurements

Figure 3.8: OEM ozone-profile retrieval on 26 July 2017 (red curve). The tropospheric tradi- tional retrieval (blue curve) extends from 2.5 km to 14 km, while the stratospheric traditional retrieval (black curve) extends from 12.5 km to 43 km. At the region where the tropospheric and stratospheric lidar measurements overlap, the OEM can smoothly makes a transition from one lidar system’s measurements to the other system’s measurements.

3.4 Tropospheric Ozone Lidar Retrievals

We have demonstrated a retrieval for stratospheric ozone profiles using the OEM and mea- surements from a stratospheric DIAL lidar Farhani et al. (2018). In addition to the combined retrieval discussed here, the OEM retrieval presented in this work can be applied separately to tropospheric lidar systems. For tropospheric DIAL lidars, the overlap function must be added to the forward model, and this is the main difference between the two systems aside from including the different parameters associated with the different choice of wavelengths in a tro- pospheric DIAL system. Another difference is that for the OHP tropospheric lidar, the analog channel counts (as discussed in details in 3.2.1) do not follow a Poisson distribution. The averaging kernels for the tropospheric retrieval is shown in Fig. 3.9. Below 14.2 km (where the horizontal dashed line is plotted), at least 90% of the ozone profile is retrieved from the measurements. Although, the retrieval extends to 16.5 km, we only consider the retrieved profile up to 14.2 km in height (above this altitude, the ozone profile starts falling into the a priori profile). In higher altitudes when the SNR drops, the averaging kernel becomes smaller and the retrieval falls back to its a priori value. The residual plots are similar to one shown in Fig. 3.2. 3.4. Tropospheric Ozone Lidar Retrievals 77

Figure 3.9: Averaging kernels for the ozone density for the measurements on 12 July 2017. The horizontal dashed line is the height cut-off above which the sensitivity of the retrieval to measurements is less than 90%. The averaging kernels are only shown every 450 m in altitude. The summation of rows in the averaging kernel matrix, for each specific height, is shown by the red curve.

The retrieval starts from 2.5 km, with 150 m steps, and extends to 14.2 km. The ozone retrieval resulting from the OEM code is plotted against the sonde profile and the traditional profiles (see Fig. 3.10). The OEM retrieval and the traditional method are within good agree- ment for most heights. Above 12 km the difference between the OEM and the traditional profile reaches to its maximum of 25%.

Figure 3.10: Both the OEM retrieval (red curve) and the traditional retrieval (blue curve) extend from 2.5 km to 14.2 km. The ozonesonde profile is plotted in green. The black dashed line defines the cut-off altitude of the retrieval.

The statistical uncertainties and vertical resolution of the traditional and OEM retrievals are shown in Fig. 3.11. The OEM retrieval has a better vertical resolution but higher uncertainty. The OEM retrieval resolution at 2.5 km is 150 m and at 14.2 km it becomes 600 m. As shown in the right panel of Fig. 3.11, the vertical resolution in the OEM retrieval is 200 m until about 78 Chapter 3. Improved ozone UTLS DIAL measurements

10 km altitude. At an altitude of 5.5 km, where the photon counting signals are added, a small spike is observed. The traditional vertical resolution starts at 150 m and reaches to 1500 m at 14.2 km. The uncertainty of the OEM and the traditional method are similar for the first few kilometers. At 5.5 km, where the digital measurements begin, both methods have an un- certainty smaller than 1%. Above 5.5 km, the uncertainty in the OEM retrieval grows larger, and at 14.2 km it becomes 10.2%. The uncertainty of the traditional retrieval becomes larger as well, however at 14.2 km it is 7%. As shown in our stratospheric retrieval, the data and retrieval grids, as well as the correlation lengths, in the OEM method can be chosen to trade off larger vertical resolution for lower statistical uncertainty, like in the traditional method.

Figure 3.11: Left panel: The statistical uncertainty of the OEM retrieval (red curve) as well as the statistical uncertainty of the traditional retrieval (blue curve) for 12 July 2017. Right panel: Vertical resolution of the OEM retrieval is shown in red, while the vertical resolution of the traditional retrieval is shown in blue. The spike at 5.5 km in the OEM uncertainty is from the inclusion of the digital channels at this height.

OEM tropospheric ozone profile have also been retrieved for measurements on the nights of 14 July 2017 and 26 July 2017. The OEM retrieval for the night of 14th July is in good agreement with the traditional method, and for most heights, the difference between the two methods is small. At 11.5 km, the OEM retrieval has a better agreement with the sonde pro- file, but the difference between the two methods is only 2.5%, which is within their statistical uncertainty (see Fig. 3.12). On 26 July, at all altitudes above 3 km, the difference between the traditional and the OEM retrievals is less than 2%. 3.5. Summary 79

Figure 3.12: Comparison of tropospheric OEM ozone DIAL retrievals for the nights of 14 July and 26 July 2017 (red curves), the traditional method (blue curves), and ozonesonde measure- ments (green curves). The horizontal dashed line is the altitude below which the OEM retrieval is mostly independent of the a priori ozone profile assumed. Left panel: 14 July 2017; right panel: 26 July 2017.

3.5 Summary

We have introduced a first-principle OEM retrieval for tropospheric ozone profiles, as well as for a combination of tropospheric and stratospheric ozone profiles. Using the DIAL lidar measurements, we retrieved ozone profiles starting in the free troposphere and extending to the upper stratosphere. The results from our implementation of the OEM are summarized below.

1. The forward model uses the lidar equation and works directly with the raw measure- ments. The forward model provides a robust estimate of ozone profiles for clear nights.

2. The combined stratospheric-tropospheric DIAL OEM retrieval calculate a single ozone profile consistent with all the measurements.

3. A new retrieval method for tropospheric DIAL ozone lidars is given in the Appendix.

4. We used four different channels for tropospheric ozone retrievals, and eight different channels for the stratospheric-tropospheric ozone retrievals. The OEM has the advantage of using all these measurements at the raw (Level 0) stage; thus, no gluing or merging of profiles is needed.

5. For the tropospheric retrievals, the traditional method and our OEM retrieval produce similar results. 80 Chapter 3. Improved ozone UTLS DIAL measurements

6. In the UTLS, the OEM retrieval smoothly transitions from one lidar system to the other system. The vertical resolution of the OEM retrievals in this region is about 600 m, and the retrieval uncertainty due to measurement noise does not exceed 7%.

7. Both tropospheric and tropospheric-stratospheric retrievals provide a full uncertainty budget which includes both statistical and systematic uncertainties.

3.6 Conclusions

Our OEM implementation brings benefits to the analysis of DIAL ozone measurements. Our retrieval has no need for “gluing” or “merging” the tropospheric and stratospheric measure- ments, as all measurements are simultaneously considered while retrieving a single ozone pro- file from multiple analog and digital channels measured by the two lidars. It provides a single ozone profile consistent with the measurements from both lidar systems, and includes the ver- tical resolution as a function of height, a detailed uncertainty budget, and averaging kernels to facilitate comparisons with other instruments. While our initial implementation of our retrieval for ozone in the free troposphere and stratosphere has advantages, it also has limitations. Our forward model has been tested un- der clear sky conditions. However, in the UTLS region, clouds and significant aerosol loads can occur. We are planning to augment our forward model to allow for inclusion of aerosols, as well as other trace gases. We have taken steps in this direction by including the retrieval of overlap and particle extinction in our forward model for rotational-Raman temperature re- trievals Mahagammulla Gamage et al. (2018). Another limitation of our current forward model is a difference in how the measurements must be handled during highly variable sky condi- tions. In the traditional method, individual scans can be corrected for dead time effects and then added in time. Our current OEM implementation assumes that scans added in times have roughly similar count rates. For a situation with high variability in count rates, such as the presence of patchy clouds, we recommend using shorter temporal integrations, retrieving the associated ozone profile, and then averaging the resulting ozone profiles. As the computational requirements for our retrievals are minimal, processing measurements in short time blocks (e.g. minutes) is practical even on a modest laptop computer. 3.6. Conclusions 81

Future work, in addition to improvements to our forward model, includes comparing our retrievals with satellite measurements in the UTLS. We also plan to re-process the OHP DIAL lidar measurements using the OEM technique.

3.6.1 Dataset Citation

G. Ancellet, ftp://ftp.cpc.ncep.noaa.gov/ndacc/meta/lidar/ga-ohp-tropo-ldr.txt 82 Chapter 3. Improved ozone UTLS DIAL measurements

Bibliography

Ancellet, G. and Beekmann, M.: Evidence for changes in the ozone concentrations in the free troposphere over southern France from 1976 to 1995, Atmos. Env., 31, 2835–2851, 1997.

Ancellet, G., Papayannis, A., Pelon, J., and Megie, G.: DIAL tropospheric ozone measurement using a Nd: YAG laser and the Raman shifting technique, JTECH, 6, 832–839, 1989.

Baray, J. L., Courcoux, Y., Keckhut, P., Portafaix, T., Tulet, P., Cammas, J. P., Hauchecorne, A., Godin Beekmann, S., Maziere,` M. D., Hermans, C., et al.: Ma¨ıdo observatory: a new high-altitude station facility at Reunion Island (21 S, 55 E) for long-term atmospheric remote sensing and in situ measurements, amt, 6, 2865–2877, 2013.

Browell, E. V.: Differential absorption lidar sensing of ozone, IEEE, 77, 419–432, 1989.

Cohen, Y., Petetin, H., Thouret, V., Marecal,´ V., Josse, B., Clark, H., Sauvage, B., Fontaine, A., Athier, G., Blot, R., et al.: Climatology and long-term evolution of ozone and carbon monoxide in the upper troposphere–lower stratosphere (UTLS) at northern midlatitudes, as seen by IAGOS from 1995 to 2013, acp, 18, 5415–5453, 2018.

Deeter, M., Edwards, D., and Gille, J.: Retrievals of carbon monoxide profiles from MOPITT observations using lognormal a priori statistics, Journal of Geophysical Research: Atmo- spheres, 112, 2007.

Eberhard, W. L.: Correct equations and common approximations for calculating Rayleigh scat- ter in pure gases and mixtures and evaluation of differences, Applied optics, 49, 1116–1130, 2010.

Eriksson, P., Jimenez, C., and Buehler, S.: Qpack, a general tool for instrument simulation and retrieval work, J. Quant. Spectrosc. Radiat. Transfer, 91, 47–64, 2005.

Fally, S., Vandaele, A. C., Carleer, M., Hermans, C., Jenouvrier, A., Merienne,´ M. F., Coquart, B., and Colin, R.: Fourier transform spectroscopy of the O2 Herzberg bands. III. Absorp- tion cross sections of the collision-induced bands and of the Herzberg continuum, J. Mol. Spectrosc., 204, 10–20, 2000. BIBLIOGRAPHY 83

Farhani, G., Sica, R. J., Godin-Beekmann, S., and Haefele, A.: Opti- mal Estimation Method Retrievals of Stratospheric Ozone Profiles from a DIAL Lidar, AMTD, 2018, 1–22, https://doi.org/10.5194/amt-2018-310, URL https://www.atmos-meas-tech-discuss.net/amt-2018-310/, 2018.

Fletcher, R.: Practical methods of optimization, John Wiley & Sons, 2013.

Forster, F., Piers, M., and Shine, K. P.: Radiative forcing and temperature trends from strato- spheric ozone changes, Journal of Geophysical Research: Atmospheres, 102, 10 841–10 855, 1997.

Gaudel, A., Ancellet, G., and Godin-Beekmann, S.: Analysis of 20 years of tropospheric ozone vertical profiles by lidar and ECC at Observatoire de Haute Provence (OHP) at 44 N, 6.7 E, Atmospheric Environment, 113, 78–89, 2015.

Godin, S., Carswell, A. I., Donovan, D. P., Claude, H., Steinbrecht, W., McDermid, I. S., McGee, T. J., Gross, M. R., Nakane, H., Daan, Swart, P. J., Bergwerff, B. B., Uchino, O., von der Gathen, P., and Neuber, R.: Ozone differential absorption lidar algorithm intercom- parison, Appl. Opt., 38, 6225–6236, 1999.

Godin-Beekmann, S., Porteneuve, J., and Garnier, A.: Systematic DIAL lidar monitoring of the stratospheric ozone vertical distribution at Observatoire de Haute-Provence (43.92[de- gree]N, 5.71[degree]E), J. Environ. Monit., 5, 57–67, 2003.

Halldorsson, T. and Langerholc, J.: Geometrical form factors for the lidar function, Appl. Opt., 17, 240–244, 1978.

Heath, D. F., Schlesinger, B. M., and Park, H.: Spectral change in the ultraviolet absorption and scattering properties of the atmosphere associated with the eruption of El Chichon:´ Strato- spheric SO2 budget and decay, Eos Trans. AGU, 64, 197, 1983.

Hedin, A. E.: Extension of the MSIS thermosphere model into the middle and lower atmo- sphere, J. Geophys. Res.: Space Physics, 96, 1159–1172, 1991. 84 Chapter 3. Improved ozone UTLS DIAL measurements

Holton, J. R., Haynes, P. H., McIntyre, M. E., Douglass, A. R., Rood, R. B., and Pfister, L.: Stratosphere-troposphere exchange, Rev. Geophys., 33, 403–439, 1995.

Hunt, W. H. and Poultney, S. K.: Testing the linearity of response of gated photomultipliers in wide dynamic range laser radar systems, IEEE Trans. Nucl. Sci, 22, 116–120, 1975.

IPCP: Climate change 2007: The physical science basis, Agenda, 6, 333, 2007.

Krueger, A. J. and Minzner, R. A.: A midlatitude ozone model for the 1976 US Standard Atmosphere, Journal of Geophysical Research: Atmospheres (19842012), 81, 44774481, 1976.

Leblanc, T., Sica, R. J., van Gijsel, J. A. E., Godin-Beekmann, S., Haefele, A., Trickl, T., Payen, G., and Liberti, G.: Proposed standardized definitions for vertical resolution and uncertainty in the NDACC lidar ozone and temperature algorithms Part 2: Ozone DIAL uncertainty budget, amt, 9, 4051–4078, 2016.

Leblanc, T., Brewer, M. A., Wang, P. S., Granados-Munoz,˜ M. J., Strawbridge, K. B., Travis, M., Firanski, B., Sullivan, J. T., McGee, T. J., Sumnicht, G. K., Twigg, L. W., Berkoff, T. A., Carrion, W., Gronoff, G., Aknan, A., Chen, G., Alvarez, R. J., Lang- ford, A. O., Senff, C. J., Kirgis, G., Johnson, M. S., Kuang, S., and Newchurch, M. J.: Validation of the TOLNet lidars: the Southern California Ozone Observation Project (SCOOP), Atmos. Meas. Tech., 11, 6137–6162, https://doi.org/10.5194/amt-11-6137-2018, URL https://www.atmos-meas-tech.net/11/6137/2018/, 2018.

Logan, J. A.: Tropospheric ozone: Seasonal behavior, trends, and anthropogenic influence, Journal of Geophysical Research: Atmospheres, 90, 10 463–10 482, 1985.

Mahagammulla Gamage, S., Sica, R. J., Haefele, A., and Martucci, G.: Retrieval of Tem- perature From a Multiple Channel Pure Rotational Raman-Scatter Lidar Using the Optimal Estimation Method, Appl. Opt., in press, 2018.

Malicet, J., Daumont, D., Charbonnier, J., Parisse, C., Chakir, A., and Brion, J.: Ozone UV spectroscopy. II. Absorption cross-sections and temperature dependence, J. ATMOS. CHEM., 21, 263–273, 1995. BIBLIOGRAPHY 85

McDermid, I. S., Godin, S. M., and Walsh, D.: Lidar measurements of stratospheric ozone and intercomparisons and validation, Appl. Opt., 29, 4914–4923, 1990.

Megie, G. J., Ancellet, G., and Pelon, J.: Lidar measurements of ozone vertical profiles, Appl. Opt, 24, 3454–3463, 1985.

Merienne,´ M. F., Jenouvrier, A., Coquart, B., Carleer, M., Fally, S., Colin, R., Vandaele, A. C., and Hermans, C.: Improved Data Set for the Herzberg Band Systems of 16O 2, J. Mol. Spectrosc., 207, 120–120, 2001.

Papayannis, A., Ancellet, G., Pelon, J., and Megie, G.: Multiwavelength lidar for ozone mea- surements in the troposphere and the lower stratosphere, Appl. Opt, 29, 467–476, 1990.

Povey, A. C., Grainger, R. G., Peters, D. M., and Agnew, J. L.: Retrieval of aerosol backscatter, extinction, and lidar ratio from Raman lidar with optimal estimation, amt, 7, 757–776, 2014.

Ramaswamy, V., Boucher, O., Haigh, J., Hauglustine, D., Haywood, J., Myhre, G., Nakajima, T., Shi, G. Y., and Solomon, S.: Radiative forcing of climate, Climate change, 349, 2001.

Rodgers, C. D.: Inverse methods for atmospheric sounding: theory and practice, vol. 2, World scientific, 2000.

Sica, R. J. and Haefele, A.: Retrieval of temperature from a multiple-channel Rayleigh-scatter lidar using an optimal estimation method, Appl. Opt, 54, 1872–1889, 2015.

Sica, R. J. and Haefele, A.: Retrieval of water vapor mixing ratio from a multiple channel Raman-scatter lidar using an optimal estimation method, Appl. Opt, 55, 763–777, 2016.

Stohl, A., Bonasoni, P., Cristofanelli, P., Collins, W., Feichter, J., Frank, A., Forster, C., Gera- sopoulos, E., Gaggeler,¨ H., James, P., et al.: Stratosphere-troposphere exchange: A review, and what we have learned from STACCATO, J. Geophys. Res.: Atmospheres, 108, 2003. Chapter 4

Machine Learning Methods in Lidar Measurements Classification

4.1 Introduction

Lidar equipment consists of a transmitter unit (laser and beam formation optics), a receiver (telescope), and a signal detector unit (photomultiplier tubes and some optics). The recorded backscattered measurements (also known as level-0 measurement scans) are co-added for a period. Before co-adding, all scans should be checked for quality purposes to remove the “bad scans” from “good scans”. Recorded measurements with extremely low power laser signals, high background counts, outliers, and scans with distorted or unusual shapes are bad scans. Over many nights, we may collect a series of scans that are not identified as bad but have lower laser power, which means they may need to be separated from high power scans before processing for geophysical quantities. Furthermore, depending on the lidar system and the purpose of the measurements, scans with traces of clouds or aerosol might be classified separately. Over a night of measurement, atmospheric changes and laser fluctuations can cause both the background counts and the signal power to change abruptly. These changes make it difficult to identify bad scans quantitatively. Furthermore, identifying outlier signals is a complex task. During an observation night, signal quality can change for different reasons including any change in light pollution, the appearance of thin clouds, and drops in laser power.

86 4.2. Instrumentation 87

Scans are often analyzed daily and classified manually as good or bad. A more common way of classifying scans is to define a threshold for the signal-to-noise ratio at some altitude: any scan which does not meet the pre-defined threshold value is flagged bad. In this method, bad scans may be incorrectly flagged as good, as they might pass the threshold criteria, but have the wrong shape at other altitudes. Recently, Wing et al. (2018) suggested that a Mann- Whitney-Wilcoxon rank-sum metric could be used to identify bad scans. In the Mann-Whitney- Wilcoxon test, the null hypothesis that the two populations are the same is tested against the alternate hypothesis that there is a significant difference between the two populations. The main advantage of this method is that it can be conducted when the data distribution is not Gaussian. However, defining a local median is a subjective task. Using state-of-the-art machine learning techniques, we have introduced an automated and robust classification method for lidar scans. We present our result for two different lidar sys- tems, the Purple Crow Lidar (PCL) system and the Meteoswiss Raman Lidar for Meteorologi- cal Observations (RALMO) system. In Section 4.2, we briefly describe the experimental setup for both PCL and RALMO. In Section 4.3, we explain how machine learning (ML) works. In Section 4.4, the algorithms used in this chapter are explained in detail. In Section 4.5, we show our results, and in Sections 4.6 and 4.7, the summary of the ML approach is provided, and is compared with other methods.

4.2 Instrumentation

The PCL is a Rayleigh-Raman lidar which has been operational since 1992. From 1992 to 2010, the lidar was located at the Delaware Observatory (42.5◦N, 81.2◦W) near London, On- tario, Canada. In 2012, the lidar was moved to the Environmental Science Western Field Station (43.1◦N, 81.3◦W). The PCL uses the second harmonic of an Nd: YAG solid-state laser. The laser operates at 532 nm and has a repetition rate of 30 Hz at 1000 mJ. The primary receiv- ing optics is a liquid mercury mirror with the diameter of 2.2 m. Currently, the PCL has four detection channels:

1. A high gain Rayleigh channel which detects the backscattered counts from 30 km to 110 km altitude. 88 Chapter 4. Machine Learning Methods in Lidar Measurements Classification

2. A low gain Rayleigh channel which detects the backscattered counts from 30 km to 110 km altitude. This channel is optimized to detect counts at lower altitudes where the high intensity backscattered counts can saturate the detector, and cause nonlinearity in the observed signal. Thus, using the low gain channel, at lower altitudes, the signal remains linear.

3. A Nitrogen Raman channel which detects the Raman shifted backscattered counts from 0.5 km to 20 km altitude.

4. A Water Vapour Raman channel which detects the Raman shifted backscattered counts from 0.5 km to 20 km altitude.

The Rayleigh channels are used for atmospheric temperature retrievals, and the water vapour and Nitrogen channels are used together to retrieve relative humidity profiles. Details about PCL instrumentation can be found in Sica et al. (1995). The RALMO system is located at the MeteoSwiss meteorological station in Payerne, Switzer- land (46.8◦N, 6.9◦E), and has been fully operational since 2008.For RALMO Nd: YAG laser produces a beam at 354.7 nm. The laser has a maximum energy of 400 mJ per pulse, and has a repetition rate of 30 Hz. To collect the backscattered photon counts, four f/3.33 parabolic mirrors, each with a diameter of 30 cm, are used. The RALMO system has the following measurement channels:

1. A Rayleigh channel which detects the backscattered counts from 0.3 km to about 30 km in altitude; this signal is used to retrieve the temperature profile.

2. A water vapour channel which is optimized to make measurements from 0.3 km to 20 km altitude during night time and from 0.3 km to 15 km altitude during day time.

3. A Nitrogen channel which is optimized to make measurements from from 0.3 km to 20 km altitude during night time and from 0.3 km to 15 km altitude during day time.

Raman scattering from the water vapour and Nitrogen channels are used to retrieve water vapour mixing ratio profiles (Cooney, 1970; Melfi, 1972). More specifications about the li- dar can be found in Dinoev et al. (2012). 4.3. A Brief Introduction to Machine Learning 89

4.3 A Brief Introduction to Machine Learning

Machine learning (ML) is a branch of artificial intelligence that uses algorithms to predict out- comes. Machine learning observations, called instances, are described by number of variables called attributes. Supervised machine learning is a branch of machine learning in which in- stances are labeled. Thus, each instance has a set of independent attributes x and a dependent attribute y which is the label for each instance. The purpose of ML is to compute a function to map x to y. Formally, we are trying to learn a prediction function f (x): x → y which minimizes the expectation of some loss function L(y, f ) = Σ(ytrue − ypredicted), where ytrue is the actual value (label) and ypredicted is the prediction generated from the prediction function. The category of modeling can be either classification or regression. In classification, the target is a discrete value, and in regression, the target is a continuous quantity. In unsupervised learning, instances are not labeled, and the purpose of the method is to find similarities within the data. Hence, clustering instances are the primary goal in the unsupervised approach. Un- supervised ML can be used to detect outliers in a dataset. Here, in regards to their application in lidar data classification, both supervised and unsupervised techniques are explained.

4.3.1 Supervised Approach

In supervised learning, the dataset is divided into a training set and a test set. The training set contains both x (attributes) and y (output values). The test set only contains x values. For our lidar scan classification, we have a training set in which, for each scan, photon counts at each altitude are considered as an attribute. The classification of the scan is the output value. Thus the training set is a matrix of size (m ∗ n), and each row of the matrix presents a lidar scan in the training set. The columns of this matrix (except the last column), are photon counts at each altitude. The last column of the matrix shows the classifications of each scan as shown in 4.1. Here, all scans with a distorted shape, low laser power, or high background counts are labeled as 0. The scans with lower than normal laser power (but still in an acceptable range) are labeled as 1. The scans with indications of clouds are labeled as 2. All other scans which are scans with high laser power and low background counts are labeled as 3. Examples of each of these scans are shown in Fig. 4.2. 90 Chapter 4. Machine Learning Methods in Lidar Measurements Classification

Figure 4.1: The training set is a matrix of size (m∗n), in which each row represents one dataset (here a lidar scan), and each element in each column (except for the n-th column) represents an attribute (for lidar scans: photon counts at each altitude). The n-th column holds the label for each scan (here we have four different labels: 0, 1, 2, and 3).

The goal of a training set in supervised learning is to train the machine (i.e., have it find a mapping function between x and y) such that it can predict the output of an unlabeled dataset correctly. Following the training phase, the test set can be used to check the accuracy of the ML model. There is no defined fraction of data that should be used for the training phase, but it is essential to select an unbiased set of training data which can represent the whole dataset. In the training phase, a common issue arises when the developed algorithm becomes too complex. Although, the algorithm provides highly accurate results for the training set (often approaching 100% accuracy) it fails to perform well in the test phase. This is known as the overfitting problem. To overcome this problem, a simpler algorithm can be used. Compared to complex algorithms, simple algorithms have a lower accuracy rate in the training set; in return, in the test phase, they can perform better. Thus, highly complex models perform well in the training phase, but they have a poor generalization. It is said that they have high variance and low bias. Although simple models cannot perform perfectly in the training phase, they have a better generalization ability. Thus, they have low variance and high bias. Hence, there is a “trade-off” between the variance and bias, such that a model cannot be complicated and perform perfectly for a set of data, and at the same time, have an excellent generalization. The best practice is to take a model with a level of complexity that shows an acceptable performance 4.3. A Brief Introduction to Machine Learning 91

Figure 4.2: Example of scans taken by PCL for Rayleigh and Raman channels. Panel (a) shows examples of bad scans. In this plot, the signals in blue and dark red have extremely low laser power, the purple signal has extremely high background counts, and the signal in orange has a distorted shape and high background counts. Panel (b) shows two examples of good signals: the signal in red has a high laser power compared to the blue curve. In the supervised approach, the red signal is labeled 3, and the blue signal is labeled 1. Panel (c) shows a scan in which clouds occur at lower altitudes.

for both the training and test sets. Furthermore, selecting a larger dataset (if possible) can help to avoid overfitting and make a better model. Finally, it is worth noting that for inverse modeling methods (like OEM) discussed in pre- vious chapters, y is considered as the observable and x is the quantity of interest. However, in machine learning, x is given and y is predicted.

4.3.2 Unsupervised Approach

In unsupervised learning, all of the given data are unlabeled. This method is mostly used to find the underlying structure of the data by clustering, then looking for similarities. In unsupervised algorithms, some similarity measures are considered for clustering. After the algorithm creates clusters from the data, the user should look at each determined cluster to confirm that data in each independent cluster are similar to each other and different from data in other clusters. Moreover, by comparing data points from different clusters, the differences between the clusters should be identified. Here, we present a simple example to demonstrate how an unsupervised algorithm can make clusters. We use the iris flowers dataset (Dheeru and Karra Taniskidou, 2017) which includes three different types of irises (Setosa, Versicolour, and Virginica). Each individual iris flower has four attributes (features); this means that the dataset 92 Chapter 4. Machine Learning Methods in Lidar Measurements Classification is in a four dimensional (4D) space. In many unsupervised learning methods, the focus is to re- duce the dimension of datasets while preserving the most important characteristics of it. Here, we reduce the dimension from 4D to a 2D space, as a 2D space is possible for us to visualize. As shown in Fig. 4.3 with the help of an unsupervised learning algorithm (in this example we used the t-distributed stochastic neighbor embedding (t-SNE) method) we can make three clus- ters. Each cluster represents a type of iris. For our lidar scan classification, we are interested in seeing if the lidar scans based on their similarities (similar features) can be clustered. For our clustering task, a good ML method will distinguish between high background counts, low laser power scans, clouds, and high laser scans, and put each of these in a different cluster. A detailed description of unsupervised learning can be found in (Hastie et al., 2009).

Figure 4.3: Using a t-SNE unsupervised algorithm, the iris dataset is clustered into three dif- ferent types. Setosa is shown in violet, versicolor is shown in blue and virginica is shown in yellow.

4.4 Learning Algorithms

Many algorithms have been developed for both supervised and unsupervised learning. Here, we introduce those which are used in this chapter. 4.4. Learning Algorithms 93

4.4.1 Support Vector Machine Algorithms

The Support Vector Machine (SVM) algorithms are popular in the remote sensing community because they can successfully handle small training sets, while producing highly accurate pre- dictions (Mantero et al., 2005; Foody and Mathur, 2004). Moreover, unlike some statistical methods such as Maximum Likelihood Estimation, which assume data is distributed normally, the SVM does not require this assumption. This property makes it suitable for datasets with unknown distributions. Here, we briefly, describe how SVM works. More details on the topic can be found in (Vapnik, 2013; Burges, 1998). The SVM algorithm finds an optimal hyperplane which separates the dataset into a distinct predefined number of classes (Vapnik, 2013). A hyperplane is a subspace with a dimension which is one less than its ambient configuration space. The optimal hyperplane is a deci- sion boundary that minimizes misclassifications and is obtained from an iterative process in the training phase. The simplest form of SVM is a linear binary classifier in which a two- dimensional input space will be divided into two-class classification (see Fig. 4.4). As shown in the figure, an optimal hyperplane provides a maximum margin (separation line between the two different classes) between the two classes. Here, the hyperplane is used to separate the star logos from the sun logos. In the figure, few objects are misclassified. It is possible to obtain more complex hyperplanes; however, it can lead to the overfitting problem. To define the hyperplane, there is no need to use all the data in the training set; and only a subset of data points that lie on the margin (called support vector) is used. To use SVM as a multi-class classifier, some adjustments need to be made to the simple SVM binary model. Methods like a directed acyclic graph, one-against-all, and one-against- others are among the most successful techniques for multi-class classification. Details about these methods can be found in Knerr et al. (1990).

4.4.2 Decision Trees Algorithms

Decision trees are nonparametric algorithms. Decision trees allow complex relations between inputs and outputs, to be modeled. Moreover, decision trees are the foundation of both ran- dom forest and boosting methods. A comprehensive introduction to the topic can be found in 94 Chapter 4. Machine Learning Methods in Lidar Measurements Classification

Figure 4.4: Example of linear SVM: the hyperplane separates sun and star logos. The subset of data points within the margin (support vector) is the most important data point as they play an important role in defining the optimum hyperplane. IN SVM a wider margin is preferable, as it divides two classes. This figure is adapted from: (Burges, 1998).

Quinlan (1986), here, we briefly describe how a decision tree is built. A decision tree is a set of (binary) decisions represented by an acyclic graph directed out- ward from a root node to each leaf (see Fig. 4.6). Each node has one parent (except the root), and can have two children. A node with no children is called a leaf. Decision trees can be complex and this depends largely on the dataset. Later, the tree is simplified by pruning (from leaves to upper parts of the tree). To grow a decision tree, the following steps should be taken.

• Defining a set of candidate splits: We should answer a question about the value of a selected input feature to split the dataset into two groups.

• Evaluating the splits: Using a score measure, at each node, we can decide what the best question is to be asked and what the best feature is to be used. As the goal of splitting is to find the purest learning subset (in each leaf, we want the output labels to be the same), the purity improves of each split candidate should yield to the score measure.Shannon 4.4. Learning Algorithms 95

Entropy (see below) is used to evaluate the purity of each subgroup. Thus, a split that reduces the entropy from one node to its descendent is favorable

• Deciding to stop splitting: We should set some rules to define when the splitting should be stopped, and a node becomes a leaf. This decision can be data-driven. For example, we can stop splitting when all objects in a node have the same label (pure node). The decision can be defined by a user as well. For example, we can limit the maximum depth of the tree (length of the path between root and a leaf).

Figure 4.5: The schematics of a decision tree is shown. From the root (a node with no parents), a decision question is asked and answered in the form of yes (1) or no (0). The answers (yes or no) generate two children. Leaves are the nodes with no children. Ideally, each leaf is pure (all the points inside it has the same classification).

In a decision tree, by performing a full scan of attribute space the optimal split (at each lo- cal node) is selected, and irrelevant attributes are discarded. This method allows us to identify the attributes that are most important in our decision-making process. In summary, the sim- plicity of decision trees makes them suitable algorithms for both classification and regression processes. The metric used to judge the quality of the tree splitting is Shannon entropy (Shannon, 1948). Shannon Entropy describes the amount of information gained with each event and is 96 Chapter 4. Machine Learning Methods in Lidar Measurements Classification calculated as follows:

H(x) = −Σpi log pi (4.1) where pi represents a set of probabilities. For example, if we toss a coin and are interested in the Shannon entropy of the output, we would calculate it as follows. The probabilities can be written as p (probability of the outcome being heads) and q = 1− p (probability of the outcome being tails). The Shannon entropy is then:

H(x) = −(pi log pi + qi log qi). (4.2)

where H(x) defines how much information is gained from the observation. To understand this result consider an unfair coin, where both sides of the coin are heads, and before doing any measurements, we know that flipping the coin will give us no new information (that is p = 1, q = 0 and H(x) = 0). In comparison, in the process of tossing a fair coin (where p = 0.5 and q = 0.5) we do not know the outcome of any individual flip, so a measurements gives us a maximum amount of information, that is H = 1.

4.4.3 Random Forests

A critical issue with trees is their high variance. This high variance can occur when a tree gets too complex in order to make pure leaves. To overcome this issue, we can grow an ensemble of trees which is known as the random forest (RF) method. If the number of trees is large, overfitting is not a problem, (Breiman, 2001). In an RF algorithm, to decide the class type all trees vote, and the vote of the majority is selected as the output. In the RF algorithm, to grow each tree, the bootstrap aggregating (bagging) method is used to sample from the training set. In bagging, a dataset is iteratively resampled with replacements. The bagging allows a data point to be included in a sample more than once. Parameters which can significantly influence RFs are the number of trees and the tree depth. As mentioned earlier, increasing the number of trees can help with the overfitting problem, and growing more trees in a forest yield a smaller prediction error. Finding the optimal depth of each tree is a crucial task. While leaves in a short tree may contain heterogeneous data (the 4.4. Learning Algorithms 97 leaves are not pure), tall trees can suffer from poor generalization. Thus, the optimal depth provides a tree with pure leaves and great generalization.

Figure 4.6: The schematic of a random forest: each tree casts a vote; based on majority vote, a data point is classified.

4.4.4 Gradient Boosting Tree Methods

Boosting methods are based on the idea that combining many “weak” approximation models (a learning algorithm which is slightly more accurate than 50%) will eventually boost the predic- tive performance (Knerr et al., 1990; Schapire, 1990). Thus, many “local rules” are combined to produce highly accurate models. In the gradient boosting method, simple parametrized models (base models) are sequen- tially fitted to current residuals (known as pseudo-residual) at each iteration. The residuals are the gradients of the loss function (they show the difference between the predicted value and the true value) which we are trying to minimize. The Gradient Boosting Trees (GBT) algo- rithm is a sequence of simple trees generated such that each successive tree is grown based on the prediction residual of the preceding tree with the goal of reducing the new residual (see Fig.4.7. This “additive weighted expansion” of trees will eventually become a strong classifier (Knerr et al., 1990). This method can be successfully used even when the relation between the instances and output values are complex. Compared to the RF model, which is based on building many independent models and combining them (using some averaging techniques), the gradient boosting method is based on building sequential models. 98 Chapter 4. Machine Learning Methods in Lidar Measurements Classification

Figure 4.7: Boosting methods are based on the idea that combining many “weak” approxima- tion models (a learning algorithm which is Using the t-SNE algorithm scans for the low-gain Rayleigh measurement channel on the night of May 15 2012 were clustered to four different groups. As shown in the figure, cluster number 2 has some outliers.

4.4.5 The t-distributed Stochastic Neighbour Embedding Method

The t-SNE method is an unsupervised ML algorithm which is based on Stochastic Neighbor Embedding (SNE). In the SNE, the data points are placed into a low-dimensional space such that the neighborhood identity of each data point is preserved. The SNE is based on finding the probability that data point (i) has data point (j) as its neighbor, which can formally be written as: 2 exp(−di, j) Pi, j = P 2 (4.3) k,i exp(−di,k)

2 where Pi, j is the probability of i selecting j as its neighbour and di, j is the squared Euclidean distance between two points in the high dimensional space, and can be written as:

2 2 || (xi − x j) || di, j = 2 (4.4) 2σi where σi is defined so that the entropy of the distribution becomes log κ, and κ is the “perplex- ity” which is set by the user, and its value determines how many neighbors will be around a selected point. 4.4. Learning Algorithms 99

The SNE tries to model each data point xi, at the higher dimension, by a point yi at a lower dimension such that the similarities in Pi, j are conserved. In this low dimensional map, we assume that the points follow a Gaussian distribution. Thus, the SNE tries to make the best match between the original distribution (pi, j) and the induced probability distribution (qi, j). This match is determined by minimizing the error between the two distributions, and the best match is developed. The induced probability is defined as:

exp(− || (y − y ) ||2) q i j i, j = P 2 (4.5) k,i exp(− || (yi − yk) || )

The t-SNE uses a similar approach but assumes a lower dimensional space which instead of being a Gaussian distribution follows Student’s t-distribution with a single degree of freedom. Thus, since a heavy-tailed distribution is used to measure similarities between the points in the lower dimension, the data points which are less similar will be located further from each other. The above approach gives t-SNE an excellent capability for visualizing data. For our unsupervised approach, we use this method. To demonstrate the difference between the Student’s t-distribution and the Gaussian distri- bution, we plot the two distributions in Fig. 4.8. Here, the x values are within 5 and -5. The Gaussian distribution with the mean at 0 and the Student’s t-distribution with the degree of freedom of 1 are generated. As is shown in the figure, the t-distribution peaks at a lower value and has a more pronounced tail.

Figure 4.8: Red curve: the Gaussian distribution for data points (extending from -5 to 5 in x- axis). The peak of the distribution is at its mean in 0. Blue curve: The Student’s t-distribution for the same data points. The distribution is heavy-tailed (comparing to the Gaussian distribu- tion). 100 Chapter 4. Machine Learning Methods in Lidar Measurements Classification

4.5 Result for supervised and unsupervised learning in PCL and RALMO system

To apply supervised learning to the PCL system, we randomly chose about 2000 scans from the low-gain and high-gain Rayleigh channels and the Nitrogen vibrational Raman channel (see Table.4.2). These measurement scans were taken from different nights, to represent different atmospheric conditions. For the low-gain and high-gain Rayleigh channels, the scans were labeled as “bad scans”, “scans with low laser power”, and “good scans”. For the Nitrogen channel we added one more label which represents scans with traces of cloud, called “cloudy” scans, where we must keep in mind by cloud we mean a substantial increase in scattering relative to a clean atmosphere, which could be caused by clouds or aerosol layers. Also, for scans of the Nitrogen channel, we did not find any “scan with low laser power”; thus for the Nitrogen channel this classification was not needed. Furthermore, in the Rayleigh channels because measurements start at 30 km, there is no need for a cloud classification. Furthermore, labeling the water vapour channel is not possible as there is no specific and well-defined pattern between different scans in this channel, so for most of the scans at this channel, we can not easily distinguish between “bad” and “good”. To overcome the overfitting issue we used the k-fold cross-validation technique in which the dataset is divided into k equal subsets. The “training” set contains k-1 of the subsets and one subset is used as a “test” set. The test set is not used in the training phase, and the accuracy result is reported based on the accuracy of the unseen test data. This process is then iteratively repeated k times, and each time a different test set is chosen. The accuracy score is the ratio of correct predictions to the total number of predictions. Thus an accuracy score of 0.85 means that our predictions are 85% accurate. Using the k-fold cross-validation method (with k = 5), we calculated the accuracy score of SVM, random forest, and gradient boosting algorithms. The results are shown in Table. 4.1. The number of scans that we used for the training at each category (“good”, ”bad”, “lower laser power”, and “cloudy”) and for each channel are shown in Table.4.2. As shown in Table. 4.1 all these algorithms perform excellently for all the channels. The only exception is the multi-class SVM for the Nitrogen channel, in which although the accuracy 4.5. Result for supervised and unsupervised learning in PCL and RALMO system 101

Channel multi class SVM RF GBT High-gain 99% 98% 99% Rayleigh Low-gain 98% 98% 98% Rayleigh Nitrogen 99% 91% 98%

Table 4.1: The success score for each of the algorithms for 2000 scans of PCL lidar.

Channel number of “good number of “scans number of “bad number of scan” with lower laser scan” “cloud” power” High-gain 172 40 131 not-applicable Rayleigh Low-gain 134 76 163 not-applicable Rayleigh Nitrogen 173 not-applicable 134 94

Table 4.2: Number of scans which used in each type. score is still high (91%), the good scans and cloud scans can get labeled wrongly. In the Nitrogen channel, the RF and the GBT can classify the lidar scans with an accuracy above 98%, and nearly all scans are labeled correctly; thus the Nitrogen channel classification is still performing well.

For the RALMO system, similar to the PCL system, random scans from the Rayleigh chan- nel and from the Nitrogen channel were selected. The labeling is binary as the scans are either labeled as cloud or clean. The Rayleigh channel in the RALMO system takes measurements from lower altitudes (as low as 50 m above the ground), so the cloudy scan label was used for the Rayleigh channel as well. Furthermore, as RALMO is operational during both days and nights, two different sets of data (one representing the day time scans and one representing the night time scans) were provided (see Table.4.5). The reason for this division is that the back- ground counts during the day time, due to the solar radiation, are much higher than at night. Thus, a scan which is labeled as a good scan during day, would be considered a bad scan at night. Using the k-fold Cross Validation technique, accuracy scores for training around 2500 scans in day time and night time are calculated and respectively shown in Table. 4.3, and Table. 4.4. Similar to the PCL system result, the accuracy scores are high. 102 Chapter 4. Machine Learning Methods in Lidar Measurements Classification

Channel multi class SVM RF GBT Rayleigh Night 99% 97% 99% Time Nitrogen Night 99% 96% 98% Time

Table 4.3: The success score for each of the algorithms for day time scans of the RALMO lidar.

Channel multi class SVM RF GBT Rayleigh Day 98% 97% 97% Time Nitrogen Day 96% 95% 98% Time

Table 4.4: The success score for each of the algorithms for day time scans of the RALMO lidar.

Channel number of “good scan” number of “cloud” Rayleigh day time 513 316 Rayleigh night time 513 316 Nitrogen day time 257 200 Nitrogen night time 257 200

Table 4.5: Number of scans used in each type; for day time and night time we used the same number of scans.

4.5.1 Unsupervised result

The TSNE algorithm allows scans for each night to be clustered. The clustering can differ from night to night (and day to day for RALMO). On nights/days where most scans look similar, fewer clusters are seen, and on other nights/days when the atmospheric or the instrument con- ditions were not stable, more clusters are generated. To demonstrate how clustering works, we present our result for May 15 2012 (a cloudy night) for data collected with the PCL low-gain Rayleigh channel, as well as, the clustering result for the Nitrogen channel. As shown in Fig. 4.9, the t-SNE algorithm generates three distinct clusters for the low-gain Rayleigh channel. Figure. 4.10 (left panel) shows all the signals for each of the clusters. The maximum number of photon counts and the value and the height of the background counts are the identifiers be- tween different clusters. Thus, cluster 3 with low background counts and high maximum counts (this indicates the power of the laser) represents a group of scans which are labeled as good scans in our supervised algorithms. Cluster 1 represents the scans with lower than normal laser 4.5. Result for supervised and unsupervised learning in PCL and RALMO system 103

powers, and clusters 2 shows scans with extremely low laser powers. To better understand the difference between these clusters, Fig.4.10 (right panel) shows the average signal within each cluster which shows the difference clearly. Furthermore, the outliers of cluster 2 (shown in black) identify the scans with extremely high background counts. This result is consistent with our supervised method, in which we had good scans (here is cluster 3), scans with lower laser power (here is cluster 1), and extremely low laser power scans (here cluster 2).

Figure 4.9: Using the t-SNE algorithm scans for the low-gain Rayleigh measurement channel on the night of May 15 2012 were clustered to four different groups. As shown in the figure, the cluster number 2 has some outliers.

Using the t-SNE, we have clustered scans for the Nitrogen channel with the data collected on 26 May 2012 at PCL system. This night was selected because the sky conditions changed from clear to cloudy and this was detected by the lidar system. The data from this night allows us to test our algorithm and determine how well it can distinguish cloudy scans from the non-cloudy scans. The result of clustering is shown in Fig.4.11 (left panel) in which two well- distinguished clusters are generated, where one cluster represents, the cloudy and the other represents the non-cloudy scans. The averaged signal for each cluster is plotted in Fig.4.11 (right panel).

Figure.4.12 shows results of clustering for few nights of measurements for different chan- nels. By looking at scans within each cluster we can find what are the similarities among scans of a cluster, and decide how to label each cluster. Thus, the t-SNE is a powerful tool which can be used to visualize our data (by reducing its dimensionality) for each night of measurements, and to cluster them. 104 Chapter 4. Machine Learning Methods in Lidar Measurements Classification

Figure 4.10: Left Panel: All scans collected for the PCL system from the low-gain Rayleigh channel which were plotted for the night of May 15 2012. The green signals have extremely low power. The red line represents all signals with lower laser signal and the blue line indicates the signals which are good scans. The black lines are signals with extremely high backgrounds. Right Panel: Each line represents an average of the signals within a cluster. The black line indicates the outliers which have extremely high background counts and are outliers belonging to cluster 3 (blue curve). The red line is the average signal for scans with lower laser power. The green line is the average signal for cluster 2. The background counts in the green line start at lower latitude of 40 km, where as for the red line the background starts at 60 km and in the blue line the background starts at 85 km.

4.6 Summary

We introduce a machine learning method to classify raw lidar measurements. We used different ML methods on elastic and inelastic measurements from both the PCL and RALMO lidar systems. The ML methods we used and our results are summarized as follows.

1. We tested different supervised ML algorithms, among which the SVM, RF, and the GBT performed better, with a success rate above 90% for both PCL and RALMO systems.

2. The t-SNE unsupervised algorithm can successfully cluster scans on nights with both consistent and varying lidar scans due to both atmospheric conditions and system align- ment/performance. For example, if during the measurements the laser power dropped or clouds became present, the t-SNE showed different clusters representing these condi- tions.

3. Unlike the traditional method of defining a fixed threshold for the background counts, in supervised ML approach the machine can distinguish high background counts by looking at the labels of the training set. In the unsupervised ML approach, by looking at the 4.6. Summary 105

Figure 4.11: Figure. 4.11 (a): Using the t-SNE algorithm, scans for the Nitrogen channel on the night of May 15 2012, collected by the PCL system, were clustered into two different groups. Figure. 4.11 (b): The red line (cluster number 2) is the average of all signals within this cluster and indicates the scans in which clouds are detectable. The blue line (cluster number 2) is the average of all signals within this cluster and indicates the clear scans (non-cloudy condition).

Figure 4.12: The result of applying the t-SNE algorithm to different channels at different nights is plotted. Each cluster contains lidar scans sharing similar features. 106 Chapter 4. Machine Learning Methods in Lidar Measurements Classification

similarities between the two scans and defining a distance scale high background counts can be grouped in one cluster.

4.7 Conclusion

We successfully implemented supervised and unsupervised ML algorithms to classify lidar measurement scans. The ML is a robust method with high accuracy which enables us to pre- cisely classify thousands of lidar scans within a short period of time. Thus, with accuracy of higher than 90% this method has a significant advantage over previous methods of classifying. For example, in the supervised ML, we train the machine by showing (labeling) different scans in different conditions. When the machine has seen enough examples of each class (which is a really small fraction of the entire data base), it can classify the un-labeled scans with no need to pre-define any condition for the system. Furthermore, in the unsupervised learning method, no labeling is needed, and the whole classification is free from subjective biases of the individual marking the scans (which for large atmospheric datasets ranging over decades is important). For example, for lidars in which observers, without using any pre-defined code identify each scan by eye, a common issue arises when an observer takes a scan as a good scan while another might classify it as a bad- scan, especially for edge cases. However, as mentioned earlier, these kind of mistakes are minimized by using an unsupervised ML approach. Our results, in this chapter, indicate that ML is a powerful technique which can be used in lidar classifications. We encourage our colleagues in the lidar community to use both super- vised and unsupervised ML algorithms for their lidar scans. To investigate rare atmospheric events, we are planning to use unsupervised ML to simultaneously analyze several nights of measurements. As mentioned in the previous chapters, using OEM requires a well-defined forward model; in contrast the ML’s focus is to learn the function which maps the state vector (x) to the mea- surements (y). Thus, we are planning to apply ML for our ozone retrievals. BIBLIOGRAPHY 107

Bibliography

Breiman, L.: Random forests, Machine learning, 45, 2001.

Burges, C. J.: A tutorial on support vector machines for pattern recognition, Data mining and knowledge discovery, 2, 121–167, 1998.

Cooney, J.: Remote measurements of atmospheric water vapor profiles using the Raman com- ponent of laser backscatter, Journal of Applied Meteorology, 9, 182–184, 1970.

Dheeru, D. and Karra Taniskidou, E.: UCI Machine Learning Repository, URL http://archive.ics.uci.edu/ml, 2017.

Dinoev, T. S., Simeonov, V. B., Arshinov, Y. F., Bobrovnikov, S. M., Ristori, P., Calpini, B., Parlange, M. B., and van den Bergh, H.: Raman Lidar for Meteorological Observations, RALMO-Part I: Instrument description, amtd, 5, 6867–6914, 2012.

Foody, G. M. and Mathur, A.: A relative evaluation of multiclass image classification by sup- port vector machines, IEEE Transactions on geoscience and remote sensing, 42, 2004.

Hastie, T., Tibshirani, R., and Friedman, J.: Unsupervised learning, in: The elements of statis- tical learning, pp. 485–585, 2009.

Knerr, S., Le,´ P., and Dreyfus, G.: Single-layer learning revisited: a stepwise procedure for building and training a neural network, in: Neurocomputing, pp. 41–50, Springer, 1990.

Mantero, P., Moser, G., and Serpico, S. B.: Partially supervised classification of remote sens- ing images through SVM-based probability density estimation, IEEE Transactions on Geo- science and Remote Sensing, 43, 2005.

Melfi, S.: Remote measurements of the atmosphere using Raman scattering, Applied Optics, 11, 1605–1610, 1972.

Quinlan, J. R.: Induction of decision trees, Machine learning, 1, 81–106, 1986.

Schapire, R. E.: The strength of weak learnability, Machine learning, 5, 197–227, 1990. 108 Chapter 4. Machine Learning Methods in Lidar Measurements Classification

Shannon, C. E.: A mathematical theory of communication, Bell system technical journal, 27, 379–423, 1948.

Sica, R., Sargoytchev, S., Argall, P. S., Borra, E. F., Girard, L., Sparrow, C. T., and Flatt, S.: Lidar measurements taken with a large-aperture liquid mirror. 1. Rayleigh-scatter system, Applied optics, 34, 1995.

Vapnik, V.: The nature of statistical learning theory, Springer science & business media, 2013.

Wing, R., Hauchecorne, A., Keckhut, P., Godin-Beekmann, S., Khaykin, S., McCullough, E. M., Mariscal, J.-F., and d’Almeida, E.: Lidar temperature series in the middle atmosphere as a reference data set. Part A: Improved retrievals and a 20 year cross-validation of two co-located French lidars, Atmospheric Measurement Techniques Discussions, 2018, 1–29, 2018. Chapter 5

Conclusions and Future Work

In this thesis, a first-principle Optimal Estimation Method (OEM) is used to retrieve strato- spheric and tropospheric ozone density profiles measured by a Differential Absorption Lidar (DIAL) system located at the Observatoire de Haute Provence (France). Although the strato- spheric and tropospheric ozone profiles are retrieved individually, the OEM is capable of using all available channels from different measurements to retrieve a single optimum ozone profile extending from the troposphere to the upper stratosphere, without any need to merge results ob- tained from different analysis routines. For tropospheric and stratospheric lidar measurements, this feature of OEM offers a significant improvement over traditional methods, as the measure- ments’ region of overlap is in the upper troposphere and the lower stratosphere (UTLS) region, and the traditional method in which the two profiles were glued together had no well-defined uncertainty. Understanding the UTLS is of vital importance because even small changes in the distribution of greenhouse gases in the UTLS can significantly affect the climate. The OEM retrieval I developed in this thesis is valid for clear conditions with low to mod- erate aerosol loading. A future plan is to add the Raman channels to our retrievals to be able to provide retrievals in the presence of strong aerosol loads, as well as thin clouds. Another future direction would be to update the OHP ozone climatology using the OEM analysis. Analyzing the entire OHP database will significantly contribute to our understanding of the ozone distribu- tion, particularly in the UTLS region. The OEM-retrieved ozone profiles have lower statistical uncertainty compared to previous lidar measurements and satellites, which in the UTLS in particular will provides better constraints on ozone models. Also, improved instrument inter-

109 110 Chapter 5. Conclusions and Future Work comparison and satellite validation will be possible in both the troposphere and stratosphere by our calculations of lidar averaging kernels. I introduced a machine learning routine in which the level 0 lidar measurements are auto- matically classified in categories corresponding to bad scans, clear sky scans, and scans with cloud or aerosol layers present. This scheme is a significant improvement for the lidar com- munity, as typically groups use either simple methods based on defining a fixed threshold at a specific height or do the classification manually, which is time-consuming. The classifier is computationally fast and is suitable even for the lidars which analyze their measurements in near “real-time.” Five methods were tested, and the random forest classifiers, the support vector machines, and the gradient decision trees methods had the highest success scores. Currently, our clas- sifying techniques cannot distinguish between clouds and aerosols loads, but they can tell if clouds and/or layers are present in a measurement. We are planning to extend our machine learning routine by adding more features to our training data to be able to distinguish between clouds and aerosol layers. I also employed the t-distributed Stochastic Neighbour Embedding (t-SNE) method which is an unsupervised algorithm. The t-SNE can successfully divide our lidar data into clusters, such that scans in each cluster share similar features. The t-SNE has potential to be in important tool, as it should be able to detect unusual events (like high loads of volcanic aerosols, fires, and meteor shower traces), in addition to evaluating scan quality without requiring training data. We are planning to employ the t-SNE to classify the entire Purple Crow Lidar system at Western to search for these unusual events. Another future direction is using machine learning technique directly for ozone retrievals, eliminating the need for a pre-defined forward model. The goal in machine learning is to find a function which maps our measurements to the state of interest, allowing retrieval of the quantity of interest. We plan to extend our machine learning system to retrieve both temperature and composition, including ozone density and possibly water vapour mixing ratio. Curriculum Vitae

Name: Ghazal Farhani

Post-Secondary The University of Western Ontario Education and Astronomy Degrees: 2011 - 2013 M.Sc.

The University of Western Ontario Atmospheric Physics 2014 - 2018 Ph.D.

Honours and Faculty of Science scholarship Awards: 2014-2016

Northern Scientific Training Program 2015-2018

Related Work Teaching Assistant Experience: The University of Western Ontario 2011 - 2018

Publications:

G. Farhani, R. J. Sica, S. Godin-Beekmann, A. Haefele, “Stratospheric Ozone Density Retrieval Using the Opti- mal Estimation Method (OEM)”, 2018, EDP science G. Farhani, R. J. Sica, S. Godin-Beekmann, A. Haefele, “Optimal Estimation Method Retrievals of Stratospheric Ozone Profiles from a DIAL Lidar”, 2018, AMTD G. Farhani, R. J. Sica, S. Godin-Beekmann, G. Ancellet, A. Haefele, “Improved ozone UTLS DIAL measurements, using an Optimal Estimation Method”, 2018, ( to be submitted at applied optics) G. Farhani, R. J. Sica, S. Godin-Beekmann, A. Haefele, “The Classification of Lidar Measurement Scans, Using Machine Learning Methods”, 2018, ( to be submitted at amt)

111 112 Chapter 5. Conclusions and Future Work

Selected Conference talks and posters:

The European Geosciences Union meeting, 2018, Austria “Stratospheric Ozone Density Retrieval Using the Op- timal Estimation Method (OEM)” The International Laser-Radar Conference, 2017, Romania “The application of the OEM on ozone retrievals of the Eureka DIAL system”